Hydrol. Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Characterising the ocean frontier: A review of marine geomorphometry

Vincent Lecours1, Margaret F. J. Dolan2, Aaron Micallef3, Vanessa L. Lucieer4 1Marine Research Lab, Department of , Memorial University of Newfoundland, St. John’s, A1B 3X9, Canada 5 2Geological Survey of Norway, P.O. Box 6315 Sluppen, 7491Trondheim, Norway 3Department of Geosciences, University of Malta, Msida, MSD 2080, Malta 4Institute for Marine and Antarctic Sciences, University of Tasmania, Hobart, 7004, Australia

Correspondence to: V. Lecours ([email protected])

10 Abstract. Geomorphometry, the science that quantitatively describes , has traditionally focused on the investigation of terrestrial landscapes. However, the dramatic increase in the availability of digital bathymetric data and the increasing ease by which geomorphometry can be investigated using Geographic Information Systems (GIS) has prompted interest in employing geomorphometric techniques to investigate the marine environment. Over the last decade, a suite of geomorphometric techniques have been applied (e.g. attributes, feature extraction, automated classification) to 15 investigate the characterisation of seabed terrain from the coastal zone to the deep sea. Geomorphometric techniques are, however, not as varied, nor as extensively applied, in marine as they are in terrestrial environments. This is at least partly due to difficulties associated with capturing, classifying, and validating terrain characteristics underwater. There is nevertheless much common ground between terrestrial and marine applications and it is important that, in developing the science and application of marine geomorphometry, we build on the lessons learned from terrestrial studies. We note, 20 however, that not all terrestrial solutions can be adopted by marine geomorphometric studies since the dynamic, four- dimensional nature of the marine environment causes its own issues, boosting the need for a dedicated scientific effort in marine geomorphometry.

This contribution offers the first comprehensive review of marine geomorphometry to date. It addresses all the five main steps of geomorphometry, from data collection to the application of terrain attributes and features. We focus on how these 25 steps are relevant to marine geomorphometry and also highlight differences from terrestrial geomorphometry. We conclude with recommendations and reflections on the future of marine geomorphometry.

Keywords: Geomorphometry, terrain analysis, marine mapping, marine geomorphology, bathymetry, GIS.

1

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

1 Introduction

Geomorphology has improved our understanding of many of the Earth’s systems and surface processes (Smith et al., 2011; Bishop et al., 2012). Morphology – and in its quantitative form – is considered the most important component of geomorphology because the interpretation of other components (e.g. genetics, chronology) often starts with the 5 characterization of morphology (Speight, 1974; Minár and Evans, 2008; Bishop et al., 2012). The shape of the landscape impacts many Earth systems at all scales. For instance, broad-scale features such as mountains and valleys may dictate weather patterns (Dimri et al., 2013), vegetation and biodiversity patterns (Anderson and Ferree, 2010), and hydrological processes (Iordanishvili, 2000), while fine-scale features such as local slope may influence soil stability (Buscarnera and Di Prisco, 2013) or influence nest-site selection by certain bird species (Whittingham et al., 2002). Overall, topography is 10 known to influence gradients in moisture, energy and nutrients across the landscape (Hengl and MacMillan, 2009).

The oceans also play a fundamental role in the Earth system at multiple scales. An important part of ocean-related work addresses the physical, chemical and biological patterns and processes of the water surface and, to a lesser extent, the water column. . By contrast, seafloor research has been less spatially extensive, except for research on global, broad-scale geomorphology and processes (e.g. Ma et al., 1998). Knowledge of seafloor topography is crucial for many questions 15 (Smith, 2004). For example, seafloor topography, or bathymetry, influences surface currents (Gille et al., 2004), near-bottom currents (White et al., 2007), and ocean mixing rates (Kunze and Llewellyn Smith, 2004). Lack of knowledge on factors influenced by bathymetry can impact modelling and forecasting of different elements, from marine species distributions (McArthur et al., 2010) to climate (Jayne et al., 2004) to paths of floating objects like marine debris (Smith and Marks, 2014).

20 It is commonly stated that the oceans are 90% unexplored (e.g. Gjerde, 2006) and that more is known about the surface of Earth’s Moon (e.g. Li et al., 2014), Mars (e.g. Carr et al., 1977; Stepinski and Vilalta, 2005), Mercury (e.g. Zuber et al., 2012) or Venus (e.g. Ford and Pettengill, 1992) than about the ocean floor (Sandwell et al., 2002; Smith, 2004; Smith and Marks, 2014). However, such statements mean little without further specification or elaboration on their real meaning. The facts are that the entire ocean floor has been mapped to a resolution of a few kilometres using satellites, which generated an 25 estimation of the underwater landscape and revealed features of the Earth’s crust beneath seafloor sediments (Smith and Sandwell, 1994; Smith, 1998). However, these coarse-resolution data are often inadequate for many scientific, economic, public safety and management purposes. Applications such as tsunami hazard assessment, submarine cable and pipeline route planning, resource exploration, habitat mapping, territorial claims, ocean circulation and climate studies, and navigation were all identified as requiring more reliable, fine-scale bathymetric data (Sandwell et al. 2002). It is therefore 30 imperative to extract the most useful information that can be retrieved from the seafloor data available at resolutions finer than 5 km.

2

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Fuelled by advancements in and Geographic Information Systems (GIS) (e.g. Grohmann, 2004), the field of geomorphometry has entered a new era in recent decades (Evans and Minár, 2011; Florinsky, 2012). Geomorphometry is defined as the science that quantitatively measures terrain morphology using geosciences, mathematics, and computer sciences (Chorley et al., 1957; Mark, 1975; Pike et al., 2009). It can be divided into two sub-fields: general geomorphometry 5 (e.g. Minár et al., 2013), and specific geomorphometry (e.g. Drăguţ and Blaschke, 2006). General geomorphometry deals with continuous surfaces in order to extract terrain attributes (e.g. slope, aspect, rugosity), while specific geomorphometry aims at characterizing or extracting discrete (Evans, 1972). Methods in geomorphometry were mainly developed and tested on artificial (e.g. Jones, 1998), terrestrial (e.g. Grohmann, 2015) and extra-terrestrial settings (e.g. Bue and Stepinski, 2006). These methods are relevant for underwater applications and have been increasingly used in the last decade 10 (Lecours et al., 2015a), but differences in the input data can sometimes produce different results than expected from land- based studies, creating the additional need for a dedicated scientific effort in marine geomorphometry . Figure 1 compares the increase in publications in both marine and terrestrial (and potentially extra-terrestrial) geomorphometry over time. The numbers illustrate that marine applications of geomorphometry are more recent and less numerous than their terrestrial counterparts. However, we note that the low number of published marine applications indicated in Fig. 1 may be biased by 15 the fact that the researched terms (i.e. geomorphometry and terrain analysis) are not always used in marine studies, even where geomorphometric techniques have been employed. For instance, in Harris and Baker (2012a), all of the 57 case studies used bathymetry, 33 of them generated slope, 23 of them measured rugosity, and 14 of them calculated a topographic position index, amongst other terrain attributes (Harris and Baker, 2012b). Despite this high use of general geomorphometry techniques, the term “geomorphometry” was not used once in the 900 pages of the volume, and “terrain analysis” was only 20 mentioned twice.

To our knowledge, no review on the state-of-the-art of marine geomorphometry has ever been written. The aims of this contribution are therefore to raise awareness of this relatively recent field and to lay the basis of marine geomorphometry practices by reviewing the relevant literature to date. As illustrated in Fig. 2, this manuscript addresses the five main steps of geomorphometry identified by Pike et al. (2009) with a focus on how these steps are relevant to marine geomorphometry and 25 different from traditional, terrestrial geomorphometry. The five steps are: sample the surface (Sect. 2), generate a surface model from the sampled heights (Sect. 3), preprocessing, i.e. correct for errors and artefacts in the surface model (Sect. 4), surface analysis, i.e. derive terrain attributes and terrain features (or objects) (Sect. 5), and apply the terrain attributes and features to a specific problem (Sect. 6). We conclude with recommendations and reflections on the future of marine geomorphometry.

3

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

2 Sampling the depth of the seafloor

For centuries, the lead line was the main instrument used to determine the depth to the seafloor, until remote sensing technologies revolutionized the way we could measure bathymetry. This section introduces the principles of four types of remote sensing technology that are currently used to collect depth information: acoustic remote sensing, satellite radar 5 altimetry, optical remote sensing, and bathymetric LiDAR (light detection and ranging). Radar altimetry and optical remote sensing both collect depth information in an indirect way, while acoustic remote sensing and bathymetric LiDAR measure the seafloor directly. Radar altimetry and acoustic remote sensing are limited to deeper waters while optical remotely sensed and bathymetric LiDAR data are limited to shallower waters, although there is some degree of overlap between the depths in which the various methods can be applied. A global scale, systematic survey could be performed with satellite-based 10 methods within a few years (Sandwell et al., 2002), compared to the estimated 600 years (Carron et al. 2001) that it would take using acoustic remote sensing technologies. The different techniques are discussed in the perspective of using the information they collect to generate Digital Terrain Models (DTM) and perform geomorphometric analyses. DTMs using bathymetric data are hereafter referred to as Digital Bathymetric Models (DBM) to distinguish them from Digital Elevation Models (DEM), a term usually reserved for terrestrial elevation data though sometimes including bathymetric data for global 15 datasets. Other techniques can be used to measure depth but are less common in the literature. For instance, ground- penetrating radar can be used to detect the thickness of the water by measuring the difference between the radar echoes from the air-water and water-seafloor interfaces (Feurer et al., 2008).

2.1 Satellite radar altimetry

In the 1970s, satellite-based radar altimeters were developed as a method to study the oceans on a global scale (Douglas et 20 al., 1987), which was a significant improvement over the extent covered by very narrow ship tracks. Radar altimeters emit microwaves that bounce on the sea surface and return to the receiver, giving the altitude of the satellite over the sea surface. The topography of the surface can then be deduced and used to derive ocean circulation patterns (Fu, 1983) or define geoid models (e.g. Fernandes et al., 2000). The geoid represents the gravitational equipotential surface of the Earth relative to the mean sea level; gravity varies in space, and the anomalies in its distribution were found to be correlated with bathymetry 25 (McKenzie and Bowin, 1976; Watts 1979). Despite initial reports stating that it was impossible to derive reliable bathymetry from satellite altimeter (Keating et al., 1984; Watts and Ribe, 1984), Dixon et al. (1983) were the first to demonstrate its feasibility using real data. Several algorithms and methods to estimate and predict bathymetry from the gravitational field have since been developed (reviewed in Calmant and Beaudry, 1996 and Sandwell and Smith, 2001). However, it remains a complex process (Calmant and Beaudry, 1996) that still requires acoustic data for calibration (Smith and Sandwell, 1997).

30 The applications of altimetry-derived bathymetric data are limited to the study of broad-scale patterns, processes and features as they only provide low resolution estimates of the bathymetry (Goff et al., 2004); ocean waves create a lot of noise that 4

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

prevents the collection of fine-resolution data (Smith, 1998), and rough seafloor and thick sediments affect data accuracy (Smith and Sandwell, 1994). Technological constraints and satellite orbits also prevent data collection close to the poles and the coastline (Sandwell et al., 2002). Some authors identified weaknesses in the method and warned that predicted depths from altimetry may not be reliable and should not be used for geodynamics studies (Smith, 1993), navigation, or 5 hazard identification (Smith and Sandwell, 1994). The main advantages of altimetry-derived bathymetry are speed of collection and uniformity of coverage (Mackenzie, 1997).

Two main altimetry-derived datasets are currently used in applications of marine geomorphometry: the General Bathymetric Chart of the Oceans (GEBCO, 2014) and the Shuttle Radar Topography Mapping 30-arc second database (SRTM30_PLUS, Becker et al., 2009). They are both free datasets that were created by filling the gaps between publicly available datasets 10 from different sources with radar altimetry (Smith and Sandwell, 1994, 1997; Becker et al., 2009). These datasets have been used for instance in habitat mapping and predictive modelling (e.g. Davies et al., 2008; Knudby et al., 2013), conservation (e.g. Ross and Howell, 2013), search and rescue operations (Smith and Marks, 2014), and geomorphology (e.g. Harris et al., 2014). Many works have found these datasets to be too coarse for their purposes (e.g. Davies et al., 2008; McNutt, 2014). For instance, Vierod et al. (2014) stated: “At present, the availability of bathymetric data at a resolution sufficient to inform 15 reliable terrain attribute predictors is a major limitation to the ability of deep-sea species distribution models to make accurate predictions of the distributions of benthic organisms.” For many applications, quality can also be just as important as resolution (c.f. Sect. 4 and Sect. 6).

2.2 Optical remote sensing

Of the four remote sensing methods presented in this section, optical passive remote sensing is the least common in the 20 marine geomorphometry literature. However, it presents a cheaper alternative to LiDAR data for collecting depth information in very shallow coastal areas (Su et al., 2014), as satellites can cover large areas in less time (Lafon et al., 2002; Wang and Philpot, 2007). Two main passive remote sensing techniques are presented: one based on the interactions of electromagnetic radiations with water and one based on principles from stereoscopy.

The ability to derive depth estimates from imagery comes from the optical Beer-Lambert law of light absorbance, which 25 describes light absorption as it goes through a transparent medium (Serway and Beichner, 1983). In water, light gets absorbed exponentially as depth increases (Lyzenga, 1978). The Beer-Lambert law allows the mathematical derivation of depth estimates from the brightness values of pixels in an image, when the absorption characteristics of an area are known (Mobley et al., 2005; Carbonneau et al., 2006). Since the absorption rate of an area is dependent on water turbidity and the characteristics of the incoming energy (e.g. the intensity, angle and wavelength of sunlight), calibration with ground-truth 30 data verifying depth-colour relationships (i.e. local light absorption characteristics) is a key step in the application of this method. However, calibration is made difficult by temporal variations in the illumination characteristics of an area

5

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

(Carbonneau et al., 2006); the calibration data would ideally need to be collected at the same time as the remotely sensed data to ensure identical environmental conditions. Since reliable calibration data are particularly challenging to obtain in marine waters (Lafon et al., 2002; Dekker et al., 2011), some methods have been proposed to estimate bathymetry without ground-truth data (e.g. Fonstad and Marcus, 2005), however these are not yet widely adopted (Feurer et al., 2008). Rather 5 than using the level of absorbed energy to derive bathymetry from imagery, some authors, e.g. Maritorena et al. (1994), have used bottom reflectance, which is the level of reflected energy.

Photogrammetry applied to pairs of stereo images can also be used to build DBMs in a similar method to the technique applied on land. Although possible (e.g. Stojic et al., 1998), through-water is challenging due to the need to correct for the air-water interface (Feurer et al., 2008). Underwater photogrammetry (i.e. active remote sensing) has, 10 however, been successfully applied at a fine scale to reconstruct the digital terrain (e.g. Johnson-Roberson et al., 2010; Kwasnitschka et al., 2013). The work by Friedman et al. (2012) is noteworthy as they derived multi-scale measures of rugosity, slope and aspect from underwater stereo image reconstructions.

All types of imagery have been used to derive marine bathymetry: hyperspectral (e.g. Mobley et al., 2005; Ma et al., 2014), multispectral (e.g. Lyzenga et al., 2006; Pacheco et al., 2015), broadband colour (e.g. Westaway et al., 2003) and grayscale 15 images (e.g. Winterbottom and Gilvear, 1997). Multispectral images enable refined depth estimates by extracting information on the bottom types from non-visible spectral bands and linking the bottom types with depth (Winterbottom and Gilvear, 1997). The red band of the electromagnetic spectrum is particularly successful in detecting depth variations (Legleiter et al., 2004; Carbonneau et al., 2006).

All methods from optical passive remote sensing are limited to shallow waters and are sensitive to errors caused by waves, 20 turbidity, sunglint from specular reflection, heterogeneous and complex seafloors, and the presence of shadow that artificially increases depth estimates (Lafon et al., 2002; Louchard et al., 2003; Holman and Haller, 2013; Eugenio et al., 2015). Some of these elements can be corrected or accounted for. For instance, Knudby et al. (2010) applied ‘deglinting’ (Hedley et al., 2005; Kay et al., 2009) and water column corrections (Lyzenga, 1978), in addition to the common geometric and atmospheric corrections (Jensen, 2005), to IKONOS satellite images. From these images a 4 m resolution DBM was 25 created and rugosity was derived. The authors indicated, however, that noise prevented the use of bathymetry at depths deeper than 15 m as rugosity values were artificially increased. That depth (15 m) is often reported as being the performance limit of optical remote sensing for bathymetry retrieval (e.g. Stumpf et al., 2003), although the practical limit will vary with local sea conditions.

2.3 Acoustic remote sensing

30 The development of acoustic technologies has fuelled marine exploration probably more than any other method, by providing reliable, continuous, and relatively high-density data at ever-decreasing price per line kilometre. Acoustic waves 6

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

are the most practical vector of information in the submarine environment; since water is denser than air, acoustic vibrations propagate through water four to five times quicker (Lurton, 2010). Sound also travels greater distances underwater as there is less attenuation in the water compared to air (Lurton, 2010).

Three tools can be used to collect depth information using sound waves: sidescan sonars (SSS), single-beam echosounders 5 (SBES), and multibeam echosounders (MBES). These tools are all active sonars (Sound Navigation and Ranging) that transmit a characteristic and controlled signal in direction of the seafloor. The two-way travel time, which is the time taken for acoustic waves to travel between the source and the seafloor and back to the source again, can be measured to estimate the range between the target and the sonar and thus deduce depth. The intensity of this return (i.e. backscatter) can also be measured to provide information on the properties of the seafloor (e.g. sediment composition). These three tools can be pole- 10 mounted on the side of vessels, or mounted on the hull of vessels, on remotely operated vehicles (ROV), on autonomous underwater vehicles (AUV), or on a towed platform.

SSS were developed for military applications in the 1940s. They provide an image of the seafloor made from backscatter measurements. Despite their ability to provide information on the topographic roughness of the seafloor, SSS cannot measure seafloor relief directly. However, similarly to what is done with interferometric synthetic aperture radar, 15 interferometry can be used to create estimates of bathymetry by combining two receiving antennas. SSS are easy to deploy and cheaper than other technologies (Harris and Baker, 2012c), but not often used to collect bathymetry as they are usually mounted on a towed platform that is difficult to accurately locate in space. Collier and Humber (2012) provide an example of the use of sidescan-derived bathymetry to identify geomorphic features on the seafloor. Some techniques from specific geomorphometry are used on backscatter data to identify specific bedforms or depositional units on the seafloor based on 20 their unique acoustic signature (Greene et al., 1999; Huvenne et al., 2005; Martorelli et al., 2012) and to detect differences in reflectivity and texture patterns on the seafloor (van Lancker et al., 2012). We also recognize the potential to generate centimetre resolution bathymetric data using modern synthetic aperture sonar systems such as the HISAS 1030 (Kongsberg Maritime, 2015) from a stable AUV platform (e.g. Thorsnes et al., in press). Synthetic aperture sonars differ from traditional SSS by utilising data from several consecutive pings to synthesize a longer sonar array capable of measuring at higher 25 resolution. There are many potential benefits to this approach although, at present, the processing of bathymetric data is very computationally demanding and therefore best suited to mapping of small areas.

SBES, or fathometers, collect both depth and backscatter data by transmitting a narrow sound beam at nadir. The mapped extent is thus limited to a single track directly below the supporting platform. Some modifications to SBES were suggested to increase their performance. For instance, the split-beam echosounder uses interferometry to improve the accuracy of the 30 data and was used to determine slope directly (Fosså et al., 2005). The dual-beam echosounder uses two beams of different aperture oriented in the same direction to locate targets more accurately. Finally, the sweep sounder is a combination of several SBES mounted on a horizontal support to increase the number of soundings. Although they remain standard for ships

7

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

navigation, SBES are less and less used for mapping purposes since MBES became more affordable, but recent applications can still be found, particularly where compiled SBES data are available; a SBES bathymetric dataset of the English Channel was used by Coggan and Diesing (2012) for the broad-scale analysis of an exposed rock ridge system, by Elvenes et al. (2014) for surficial sediment and habitat mapping, and by James et al. (2012) to identify geomorphic features in a paleo- 5 valley.

MBES provide a relatively fast, high-resolution and wide coverage measurement of the seafloor, both in terms of depth and backscatter. They sweep a large swath of the seafloor by emitting a fan of narrow sound beams, and are currently the most efficient and accurate tool available to collect bathymetric and backscatter data (Costa et al., 2009; Schimel et al., 2010a, 2010b). In recent years, advancement in MBES technology has further enhanced a valuable source of seafloor data. These 10 advances have come out of traditional user groups extending the application of the data to meet new requirements and from the motivation of new user groups wanting to employ the technology. This wide ranging and ever growing community of MBES users are adapting and extending the potential of MBES data to address unique applications. MBES users have traditionally included hydrographers, navigators, engineers, marine geologists and military planners, but now we see the extension of the technology to meet the needs of maritime explorers, archaeologists, fisheries biologists, geomorphologists 15 and ecosystem modellers, to name a few. MBES is currently the main source of bathymetric data for applications of marine geomorphometry.

2.4 Bathymetric LiDAR

Bathymetric LiDAR is an adaptation of the more traditional airborne topographic LiDAR (Irish and Lillycrop, 1999; Guenther et al., 2002) and has become increasingly common in the literature in the last two decades (Brock and Purkis, 20 2009). The main difference between the two types of systems is the wavelengths used; the laser from bathymetric LiDAR uses a wavelength in the green spectrum, compared to the red/infrared wavelength of the topographic LiDAR. Recent systems, sometimes called topo-bathy LiDAR, are multispectral LiDAR and combine both types of laser, which enables the surveying of both land and water in one flight; when flying over the water, the green laser penetrates the sea surface and collects information on the water column and the seafloor while the red/infrared laser collects information on the sea surface. 25 LiDAR can also collect intensity values that, like acoustic backscatter, provide information on the characteristics of the seafloor (Costa et al., 2009).

Bathymetric LiDAR is the only technique that can collect high-resolution data in very shallow waters, which makes it especially relevant for coastal applications requiring fine-scale data (<1 m resolution) (Brock and Purkis, 2009). The efficiency of bathymetric LiDAR systems is greatly limited by turbidity, wave action, depth (up to 50-70 m in very good 30 conditions), steep slopes, and rocky substrate (Costa et al., 2009; Chust et al., 2010; Jalali et al., 2015). Current geomorphometric applications on bathymetric LiDAR data are mainly related to the exploration of coastal ecosystems (e.g.

8

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Wedding et al., 2008; Zavalas et al., 2014) and geomorphology (e.g. Arifin and Kennedy, 2011; Kennedy et al., 2014), but are likely to extend to other applications such as marine archaeology and natural hazards assessment (e.g. Solsten and Aitken, 2006). In 2015, LiDAR data represented 4.5% of the coastal data collected for the Continually Updated Shoreline Product (CUSP) compiled by the National Oceanic and Atmospheric Administration (NOAA) and the National Geodetic 5 Survey (NGS) of the United States (Graham et al., 2015).

3 Generating a surface model from sampled depths

By nature, geomorphometric analyses necessitate spatially continuous data, but not all remote sensing techniques used to collect depth samples create continuous surfaces. For instance, bathymetric LiDAR and MBES collect point data that need to be cleaned and then interpolated in order to create a full coverage, and SBES collect data in narrow lines that sometimes 10 need to be interpolated to fill in between the survey lines. This section describes the different interpolators that are available for DBM generation and the question of spatial scale of DBMs. A detailed account of the various approaches to processing (i.e. georeferencing and applying system and environment related corrections) and cleaning of data (i.e. removal of spurious depth soundings) are beyond the scope of this paper and are specific to the sensors used for data acquisition as well as industry and application-related standard practices. We do, however, touch briefly on data cleaning in Sect. 3.2 where we 15 present a method where uncertainty algorithms are used to aid data cleaning, and where the interpolation of data is intrinsically linked to the calculation of uncertainty of the bathymetric surface.

3.1 Interpolation

Hengl and Evans (2009) identified several techniques used to generate gridded DTMs from height samples for geomorphometric purposes, including inverse distance, minimum curvature, spline, kriging, polynomial regression, moving 20 average, and many others. The same methods can all be used to generate DBMs from depth samples. For instance, Ezhova et al. (2012) created a DBM from SBES data using the natural neighbour interpolation method, and Ramillien and Cazenave (1997) combined altimetry and ship-based data into a single DBM using bilinear interpolation. More rarely, triangulated irregular networks (TIN) are created from depth samples; for example, Heyman and Kobara (2012) generated a TIN from SBES data, and Foster et al. (2009) computed TINs from SBES and bathymetric LiDAR from which volumetric attributes 25 were computed.

The choice of interpolator varies depending on the type of data and the spatial arrangement of the depth samples. For instance, MBES or LiDAR data can collect very dense point clouds that require little interpolation between points, resulting in limited interpolator influence in the final DBM. On the other hand, creating a DBM for a big area from SBES data requires more interpolation as SBES only sample very narrow tracks and have a high density of points along the survey line

9

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

but no data between the survey lines. This has implications for geomorphometry as the interpolated DBM may miss important geomorphological features (depending on the distance between the survey lines), similarly to what happens with the interpolation of contour lines (Wise, 1998).

There are no optimal interpolation methods (Li et al., 2005), and it is well known that each technique has different sensitivity 5 to errors and sample distribution and that the quality of DTMs can be improved when making the appropriate choice of interpolator (Carrara et al., 1997; Hengl and Evans, 2009). By being different in nature, sampled depths may not require the same characteristics from an interpolator than sampled elevations; for instance, DBMs do not need to be hydrologically corrected as drainage analyses are futile underwater. This is why techniques were developed in recent years to address the particular characteristics of depth sampling. Here we examine on such technique, the CUBE (Combined Uncertainty and 10 Bathymetric Estimator) algorithm (Calder, 2003), which accounts for different errors specific to acoustic remote sensing (e.g. geometric and acoustic) and is incorporated in several of the most widely used bathymetric processing software used by the hydrographic survey industry and scientific community. Although not yet universally accepted as data cleaning method by the hydrographic survey industry and hydrographic agencies, who have a particular need to preserve shoal soundings and comply with strict quality control procedures to ensure safety of navigation, CUBE is widely used and of special interest to 15 more applied bathymetric data users and the related scientific community. According to Schimel et al. (2010b), CUBE could be more appropriate than traditional gridding methods to compute precise bathymetry and associated terrain attributes. CUBE is based on the spatially explicit quantification of the total propagated uncertainty (TPU) for each data point (Calder and Mayer, 2003), enabling the rejection of samples that are outside a certain uncertainty confidence level (e.g. 95% for Calvert et al., 2014). When creating the DBM, the algorithm provides vertical error estimates and statistically assigns, to 20 each pixel, the most likely depth value based on the uncertainty of each sounding within the pixel (see Dolan and Lucieer, 2014). In several bathymetric processing software offering CUBE, users can visualize not only the most probable depth for each pixel, but also the subsequent most probable depths (e.g. second or third most likely) and select the one they think is the most appropriate based on their knowledge of an area. For example, this can allow correction for occurrences when the sonar detects fish close to the seafloor instead of the seafloor itself and data cleaning did not appropriately remove these soundings. 25 Figure 3 shows some of the information that can be extracted and visualised from the application of the CUBE algorithm. When interpolating the soundings to create a DBM, the bathymetry and the horizontal and vertical components of uncertainty can be stored in a BASE (Bathymetry Associated with Statistical Errors) surface. The BASE format allows multi-attributes surface models. CUBE’s main inconvenience is that it requires a lot of ancillary data to be collected in order to compute TPU, but it is very reliable in defining the spatial pattern of errors, their importance, and helping to identify their 30 sources (Passalacqua et al., 2015). In addition to the bathymetry, a map of uncertainty can be computed, which can become very important when making decisions using the bathymetry and for onward geomorphometric analysis.

10

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

3.2 Spatial scale

Spatial scale is a term used to summarise data resolution and extent and is an important component of the interpolation of depth data to generate DBMs. Both resolution and extent of a given DBM are dictated by the survey methodology or methodologies. First, the fundamental technical limitations of the remote sensing technique that is used to collect the 5 bathymetric data will define the scale (resolution) of the DBM (Kenny et al., 2003; Van Rein et al., 2009). For instance, principles of radar altimetry limits DBMs generated with these data to coarse, usually kilometre-scale resolution, while other methods can achieve up to centimetre-scale resolution data (c.f. Sect. 2). Second, the platform that supports the system may also influence the scale (extent) of the final DBM; a remotely sensed image collected from a satellite will usually have a coarser resolution and cover a larger area than an image collected from an aircraft or an unmanned aerial vehicle.

10 Unlike systems used in optical remote sensing, radar altimetry and bathymetric LiDAR, acoustic systems do not sample the seafloor uniformly, which influences the spatial scale of the resulting DBM. The sampling density of these systems is directly dependent on depth, or more specifically on the sensor-to-seafloor distance (Lecours and Devillers, 2015). As this distance increases, the system’s footprint increases, leading to a lower sampling density for a greater area sampled at a coarser resolution. Since the seafloor is rarely perfectly flat and at a constant depth, the sampling density is almost never 15 uniform across survey areas, which can make it challenging to determine the appropriate spatial resolution of DBMs for interpolation.

Ultimately, the spatial scale of a DBM will be dictated by its intended use (see Sect. 6) which then influences the choice of the data collection method, typically following hydrographic standards (IHO, 2008) which ensure the appropriate data are acquired to ensure safety of navigation. Besides DBMs created directly from one source of survey data, we are increasingly 20 seeing DBMs generated, or pooled together from several surveys and/or sensor technologies (e.g. EMODnet, 2015). These datasets can be a valuable resource but impose additional challenges for DBM creation and geomorphometric analysis.

4 Correcting errors and artefacts in digital bathymetric models

In terrestrial applications, it is well known that all DEMs, regardless of the techniques used to collect and generate data, are influenced by uncertainty and errors (Fisher and Tate, 2006; Gessler et al., 2009). This is also true for marine applications, 25 but the properties and dynamic nature of the ocean makes DBMs more prone to errors and artefacts than DEMs (Hughes Clarke et al., 1996). As illustrated in Fig. 4, this has significant implications for marine geomorphometry, which shows as widely recognized in the terrestrial literature (Florinsky, 1998; Zhou and Liu, 2004; Oksanen and Sarjakoski, 2005) that errors and artefacts in a DTM propagate to terrain attributes and may be amplified in subsequent analyses.

11

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

As with DEMs (Harrison et al., 2009; Sofia et al., 2013), errors and artefacts in DBMs can be caused by the interpolation method (Erikstad et al., 2013), movement and positioning of the supporting platform (Hughes Clarke et al., 1996), and a temporal (Lecours and Devillers, 2015) or spatial (Hughes Clarke, 2003a, 2003b) misalignment between the different elements of the surveying system. Data from radar altimetry are the least sensitive to platform motion (Smith and Sandwell, 5 1994). However, large artefacts resulting from fine-scale noise in the gravity field (Goff et al., 2004) and the algorithms used to convert gravity data into bathymetric estimates (Dixon et al. 1983; Calmant and Beaudry, 1996) are often characteristic of these data. Similar large linear artefacts can sometimes be found in satellite images (e.g. Klemas, 2011). The level of error in bathymetric data from optical remote sensing is known to directly depend on water depth as a result of light attenuation in the water column (e.g. Liceaga-Correa and Euan-Avila, 2002). Recent studies (Leon et al., 2013; Hamylton et al., 2015) have 10 demonstrated that the integration of the spatial structure of errors improves bathymetric estimates derived from satellite images. Data collected with acoustic methods are the most susceptible to artefacts for several reasons. First, they are collected from surface vessels/platforms or underwater vehicles that can be strongly affected by environmental conditions such as waves and wind. Furthermore acoustic waves need to be corrected for sound velocity. Without this correction the data will exhibit artefacts broadly similar to those caused by an inappropriate correction of the atmospheric conditions in 15 optical remote sensing (Li & Goldstein, 1990). Sound velocity varies with temperature, salinity and pressure and the failure to correct for these variations can induce refraction artefacts in the DBM (Yang et al., 2007). This is particularly challenging as water column properties vary both spatially and temporally, especially in the coastal zone where there is the additional complication of freshwater input from rivers, and are less predictable than atmospheric conditions (Cushman-Roisin and Beckers, 2011). Tidal corrections are generally applied using data from locally installed tide gauges, or modelled tides, 20 depending on the accuracy required. Finally, since the surveying system is underwater, direct positioning using the Global Positioning System (GPS) is not possible (Roman and Singh, 2006). The level of error in the data is thus strongly influenced by the accuracy of the different instruments that provide ancillary data to estimate the position of the system’s underwater components; the positioning error is known to directly increase with depth (Rattray et al., 2014; Lecours and Devillers, 2015).

25 Figure 4 illustrates different types of errors and artefacts that can be found in bathymetric data of different types and at different scales, and their propagation to derived terrain attributes. Artefacts commonly found in bathymetric data and that often cannot be corrected using existing methods include gridding and interpolation artefacts (e.g. in the top panels), motion artefacts (e.g. middle and bottom panels), refraction artefacts, and artefacts caused by the temporal or spatial uncertainty associated with ancillary data (e.g. bottom panels). Common errors include spurious soundings (e.g. bottom panels.)

30 Artefacts in DBMs are difficult to handle properly as they cannot generally be ground-truthed (Li and Wu, 2006). Most marine environments are not easy to access and the collection of ground-truth data is often limited by technological and logistical constraints (Solan et al., 2003; Robinson et al., 2011). Consequently, ground-truthing of DBMs is not standard practice. We note however that ground-truthing is often performed for backscatter data to attempt matching sediment types 12

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

with acoustic reflectivity characteristics. As illustrated in Fig. 4, artefacts in DBMs may be present at all scales and persist, or are sometimes amplified in derived terrain attributes. For instance, artefacts in the GEBCO are common (Lecours et al., 2013; GEBCO, 2014; Fig. 4), arising mostly from the merging of datasets of different quality. When the artefacts are too big they cannot be removed with traditional filtering methods (e.g. Gaussian filtering) as this considerably affects the overall 5 quality of the surface (Passalacqua et al., 2015), and the artefacts are also difficult to overcome when deriving terrain attributes even by using multi-scale methods (Sect. 5.1). At a finer scale, Yang et al. (2007) developed an algorithm to correct refraction artefacts, although this was only partially successful. When the artefacts are smaller, it can be difficult to distinguish them from real fine-scale features such as sandwaves or iceberg scourings (Hughes Clarke et al., 1996), especially when no underwater video data are available to confirm the geomorphology of an area. This is particularly 10 challenging for marine geomorphometry as analyses are likely to capture both the real features and the artefacts (Wilson et al., 2007). Currently, the main ways to address artefacts in DBMs are to apply filtering techniques, resample the data to coarser resolutions, manually correct the data based on visual interpretation, and to use algorithms like CUBE that account for errors and uncertainty. Most marine geomorphometry applications simply disregard the presence of the remaining artefacts, excluding them for practical purposes by expert judgement.

15 5 Deriving terrain attributes and terrain features

Bathymetric data, particularly full coverage multibeam, or LiDAR data, are well suited for the generation of quantitative terrain attributes and terrain features. These attributes and feature classifications can be very useful in describing, interpreting and classifying geomorphology in the marine environment, just as their terrestrial equivalents are on land. These derived datasets can also be of further use in geological interpretation and habitat mapping and modelling. With bathymetric 20 data now available in many areas at comparable resolutions to terrestrial DEMs, depending on the survey equipment used (c.f. Sect. 2), we can extract a similar level of information to that obtainable from terrestrial DEMs. Elsewhere, global (e.g. GEBCO, 2014) and regional (e.g. IBCAO (Jakobsson et al., 2012); EMODnet, 2015) bathymetric datasets combining information from many sources have become an impressive resource and are being used routinely for many marine science applications, not least those including high seas areas which, as yet, have little detailed coverage. This section reviews both 25 the use of general (i.e. terrain attributes) and specific (i.e. terrain features) marine geomorphometry.

5.1 General geomorphometry (terrain attributes)

The calculation of terrain attributes (synonymous with terrain/topographic variables) first requires some method for mathematically representing the bathymetric surface. This surface representation is then used to calculate the required terrain attribute, and is typically achieved by either using neighbourhood analysis of raster pixels, or by fitting a polynomial 30 expression to describe the surface. A review of terrain attributes was provided by Wilson et al. (2007) in the context of 13

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

marine benthic habitat mapping and updated by Dolan et al. (2012). In addition, Brown et al. (2011) offer a useful summary of the extent to which many of these various terrain attributes have been employed within published habitat mapping studies in the period 2000 to 2011. The computations performed on DBMs, and the range of applications that these derived terrain attributes are used for, are common to many of those performed on DEMs for terrestrial applications. The key quantitative 5 difference in the underlying data is that DBMs have negative values and this has led to modification of some terrain attributes to suit bathymetric data. Besides this, differences in analysis of bathymetric DBMs versus DEMs are often more related to the meaning or application of the information from the analysis. For instance, deriving a watershed network underwater may be useful e.g. for delineating potential sediment pathways on the continental slope, but is a deviation from the original intended purpose.

10 Terrain variables can be grouped into four main types describing different properties of the terrain – slope, orientation, curvature/relative position, and terrain variability (Fig. 5). It is beyond the scope of this paper to provide details on all the various options for computation, however, we provide an overview of some of the most commonly used terrain attributes in marine-based studies, as well as an indication of some common calculation approaches (Tables

Table 1). Here we note the geomorphological relevance and ecological relevance of the various types of terrain attributes in 15 the context of seabed mapping. Whilst the effects on geomorphology are more direct, the popularity of terrain attributes in benthic habitat mapping is, to a large extent, due to their function as a surrogate (or proxy) in explaining the distribution of benthic fauna. In the absence of better, or alternate information (e.g. from high resolution oceanographic data), proxy information such as whether a given location is sheltered or exposed to dominant currents as indicate by its position relative to neighbouring terrain, can be useful in determining suitable habitat for a given species or community. An elevated position 20 for example may be advantageous for suspension feeding organisms and act as a surrogate for the direct need for food supply. Other terrain attributes may capture a proxy for shelter or other ecological advantage. This topic is discussed further by Lecours et al. (2015b) including the all-important effect of scale which is linked both to data resolution and the scale at which geomorphometric analysis is conducted.

For GIS based calculation of terrain attributes, extending the analysis window beyond the basic 3 x 3 neighbourhood is 25 particularly useful in benthic habitat mapping as it facilitates the identification of spatial scales that are relevant to the benthic communities and may also help to overcome artefacts in the DBM (Wilson et al., 2007). The multi-scale analysis methods developed by Wood (1996), which built on the work of Evans (1972, 1980), have been fundamental in establishing an appreciation of scale in marine and terrestrial geomorphometry alike. The associated software package Landserf (Wood, 2009) puts multi-scale analysis within easy reach of marine scientists and the use of Landserf for DBM analysis took off 30 following the early applications of the software to bathymetric data (e.g. Wilson et al., 2007). Although Landserf 2.3 is still used by many scientists requiring a standalone programme for geomorphometric analysis, Wood’s algorithms are now perhaps more widely used among the marine community through the GRASS module r.param.scale. The newly released

14

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

ArcGeomorphometry toolbox (Rigol-Sanchez et al., 2015) offers a means to access the Wood-Evans (and other) algorithms for geomophometric analysis, this has the potential to provide a long awaited, convenient multi-scale analysis option for ArcGIS users.

One terrain attribute that is specifically tailored to analysis of bathymetric data is the bathymetric position index (BPI) 5 (Lundblad et al., 2006), which is an adaptation of Weiss’ (2001) topographic position index (TPI) and a useful measure of relative position that is simple to calculate over different neighbourhood sizes. Although a relatively simple algorithm to implement (Lunblad’s BPI indices can be performed through the raster calculator (e.g. Wilson et al., 2007) or scripting), many marine scientists make use of the Benthic Terrain Modeler (BTM) Toolbox, which was first developed following Lundblad’s (2006) study. The current version of BTM (Wright et al., 2012) for ArcGIS 10.1 and later has seen around 4000 10 downloads in the period 2012-2015 and this figure gives a conservative estimate of how many scientists are actually using the tool (S. Walbridge, ESRI, pers. comm.). The BTM toolbox relies on ArcGIS Spatial Analyst and includes tools for calculating slope, aspect and terrain variability (rugosity, VRM) as well as methods for combining these into geomorphic zones. It was launched for the scientific community at a time when multibeam data was becoming widely available and modern marine geomorphometry was becoming established. The BTM toolbox quickly became popular as a one-stop shop 15 for terrain analysis and classification of bathymetric data, offering a slightly more tailored solution, and the ability to handle larger datasets, than Landserf, with at least the BPI index being computable at different scales ever since the first release (now joined by VRM). The utility of the BTM tool has been augmented in recent years through updating of the terrain variability and aspect indices, and by providing the tools as both an AddIn and as a standalone ArcToolbox, providing greater flexibility to users who may wish to benefit from all, or just part of, the functionality.

20 Several bathymetric data processing software (e.g. CARIS HIPS and SIPS, QPS-Fledermaus) also have built in tools for calculation of basic indices such as slope and rugosity, bringing the functions directly to the bathymetric data user and removing the need to search for and select from the vast array of available methods. This has advantages of convenience for some bathymetric data users, but in most applied projects the computation of terrain attributes and further analysis will be conducted in some generic GIS software. Although many of the commercial software are currently limited to single scale 25 analysis (3 x 3 rectangular neighbourhood) it has become easier to find tools for multi-scale analysis, either directly in open source software (e.g. GRASS), through additional toolboxes (e.g. SEXTANTE for QGIS), or via scripting. Many of these also give alternative choices for computation algorithms, the effects of which are investigated by Dolan and Lucieer (2014) using slope as an example.

Terrain variability has been a particularly popular terrain attribute in relation to benthic habitat mapping. This is largely due 30 to the generally accepted link of terrain variability with biodiversity, which has, however, not yet been fully established with regard to spatial scale (Lecours et al., 2015b). Several measures of rugosity have been applied to DBMs with some proving suitable for multi-scale analysis and others becoming problematic at larger analysis scales (Wilson, 2006). A rugosity index

15

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

comparing surface to planar area (Jenness, 2004) remains perhaps the most widely applied method in marine studies and this was implemented in early releases of BTM. Both the vector ruggedness measure (VRM; Sappington et al., 2007) now incorporated in BTM, and the more recent Arc_Chord Rugosity measure (Du Preez et al., 2014) offers alternatives that are better decoupled from slope. Where slope and a terrain variability measures are to be used in further analysis e.g. as 5 predictor variables for habitat modelling, it is particularly important that the user is aware of any autocorrelation or covariation between them, so they can be handled appropriately. Du Preez et al. (2014) lists several marine studies among those who have ignored the need for decoupling. However, with methods like VRM and Arc-Chord rugosity or toolboxes like BTM and TASSE (Lecours, 2015) now readily available we trust that future studies will make a conscious choice of the best geomorphometric analysis for their particular application.

10 5.2 Specific geomorphometry (terrain features/objects)

Compared to general geomorphometry and the use of terrain attributes, applications of specific geomorphometry are still relatively rare in the marine environment. Calculation of terrain features generally relies on the combined properties of several terrain attributes. For instance, Lecours et al. (2013) used Troeh’s classification (Shary et al., 2005), which uses different types of curvatures to identify zones of relative deflection or accumulation and transit zones, on bathymetric 15 data. The authors also adapted Weiss’ (2001) landform classification, which combines slope with TPI measures at different scales to identify up to 16 landform classes, for application within the marine environment using BPI measures.

Terrain features such as crests, troughs etc. can be extracted through the use of pixel based analysis (e.g. Blaszczynski, 1997; Wood and Dragicevic, 2007), but recently object-oriented methods for landform classification have become increasingly popular and are beginning to make their mark on marine studies (e.g. Lawrence et al., 2015) driven by an opportunity to 20 analyse the DBM in conjunction with acoustic backscatter data (an indicator of seabed sediment type) rather than analysing the DBM alone, which offers several advantages for seabed classification. Geographic object based image analysis (Geobia, Obia) has been gaining some traction in the seabed mapping community as the spatial resolution of acoustic backscatter data improves (Diesing et al., 2014). The basic processing units in object based image analysis are objects which are represented by textural changes in the acoustic backscatter image and are constrained by derived topographic variables (Benz et al., 25 2004). Geobia allows for the quantitative extraction of image textures and features to be identified in the backscatter data and the ability to relate these spatially to topographic variability (Costa and Battista, 2013). Multi-resolution segmentation is one of the most popular segmentation algorithms to delineate homogeneous seabed segments (Lucieer, 2008; Lucieer and Lamarche, 2011; Hasan, 2012; Eisank et al., 2014) and in the terrestrial literature stands out as the most successful method to delineate homogeneous terrain segments rather than landforms per se (e.g. Drăguţ and Blaschke, 2006; Drăguţ et al., 2011; 30 Blaschke et al., 2014). This has been successfully demonstrated by Ismail et al. (2015) to identify and classify submarine canyons. By combining both the spatial derivatives of the DBM with Geobia variables, the authors were able to perform an automated multiple scale terrain analysis to discriminate local and broad scale geomorphic features in the marine landscape. 16

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

This information was used not only to delineate geomorphic seafloor features but also to identify properties that might influence biodiversity in a complex terrain. Specific geomorphometry is currently not used to its full potential in the marine environment.

6 Applications of marine geomorphometry

5 This section introduces some of the most common applications of geomorphometry in the marine environment: habitat mapping, marine geomorphology and geohazards, change detection, and hydrodynamics and modelling. Other applications of marine geomorphometry can also be found in the literature (e.g. underwater archaeology; Passaro et al., 2013) but are not discussed here. Figure 6 illustrates the proportion of the marine studies represented in the lower graph of Fig. 1 that could be classified in any of the types of applications that are presented in this section.

10 6.1 Marine habitat mapping

Benthic habitat mapping is one of the major applications areas where the use of marine geomorphometry has grown in recent years. Linked to the rise in the use of multibeam data for benthic habitat mapping (Brown et al., 2011) the vast majority of habitat mapping studies with access to good bathymetry data are now using, or at least testing, some form of terrain attribute or feature classification in their habitat mapping activities, even though we note that many of these are not yet reflected in 15 the peer-reviewed literature. Among the habitat mapping community several approaches to habitat mapping are common, many of which directly incorporate biological data, such as modelling species (e.g. Davies et al., 2008) or biotope distributions (e.g. Elvenes et al., 2014) and others which are primarily based on physical attributes deemed relevant for the distribution of benthic fauna (e.g. Micallef et al., 2012; Ismail et al., 2015). Geomorphometry is equally useful for both these approaches and those that combine both aspects (e.g. Tempera et al., 2012) and this discussion is relevant to an all- 20 encompassing definition of habitat mapping (Lecours et al., 2015b). Figure 7 illustrates how terrain attributes are typically used in the production of predictive seabed habitat maps, providing an invaluable suite of full coverage predictor variables which are used together with point samples of observed habitat as the input data to modelling.

Harris and Baker (2012b) provide a summary of surrogate variables used for habitat mapping studies in the volume ‘Seafloor Geomorphology as Benthic Habitat: GeoHAB Atlas of Seafloor Geomorphic Features and Benthic ’ including many 25 terrain attributes that have been applied across a multitude of approaches to habitat mapping worldwide. The issue of surrogacy is also discussed in this volume as well as by Lecours et al. (2015b) and McArthur et al. (2010). The case studies presented in the GeoHAB Atlas, and other published studies, vary in the degree to which they have established the ecological relevance of the terrain attributes and/or feature classifications used. For geomorphological variables to really be useful predictors of seafloor habitat, the relationship between habitat and specific variables first needs to be established.

17

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Apart from depth, which all of the geomorphological variables are derived from, different shapes or attributes of the seafloor will be relevant to different species at different scales over different bathymetric and biogeographic zones. Bathymetry is known to be a first order influence on species distribution. There are a number of recent papers describing the potential of terrain attributes to act as surrogates of species distribution (Lucieer et al., 2013; Hill et al., 2014). The relationships are 5 validated using several different statistical methods that either test terrain attributes against biological or ecological data, or combine terrain attributes with other environmental data and perform classifications to differentiate between the different habitats (Thiers et al., 2014).

In an example by Rengstorf et al. (2013), habitat suitability models for the cold-water coral Lophelia pertusa were developed based on full coverage multibeam bathymetry on the Irish continental margin. Maximum entropy modelling was used to 10 predict L. pertusa reef distribution at a spatial resolution of 0.002° (250 m). Coral occurrences were assembled from public databases, publications and video footage, and filtered for quality. Environmental predictor variables were produced by re- sampling of global oceanographic data sets and a regional ocean circulation model. Multi-scale terrain parameters were computed from multibeam bathymetry at 50 m resolution. In a related study, Rengstorf et al. (2012) examined the effect of bathymetric data resolution on terrain attributes used to predict coral distribution, resampling the original 50 m resolution 15 bathymetry from the Irish National Seabed Survey at successively coarser intervals up to 1 km. They concluded that terrain attributes derived from higher resolution bathymetry are required to adequately detect the topographic features relevant to corals. In a further related study Rengstorf et al. (2014) examined the relative importance of terrain attributes and hydrodynamic variables (e.g. current speed, vertical flow, temperature etc.) on models of cold-water coral distribution, concluding that combining the environmental information from these two sources leads to improved predictions over the 20 spatial scales in question.

At a much finer resolution (~1 m) species−habitat relationships were examined across a marine reserve on the south-eastern coast of Tasmania using boosted regression tree analyses (Cameron et al., 2014). The most important explanatory variables of community diversity were those describing the degree of reef aspect deviation from east and south (seemingly as a proxy for swell exposure), reef bathymetry (depth), low rugosity and slope. These models could account for up to 30% of the 25 spatial variability in measures of species diversity. As biological data at scales relevant to acoustic or remote sensing data such as that from AUVs, ROVs and diver surveys become available on national or international databases such as the Census for Marine Life and the Ocean Biogeographic Information System (OBIS) the ability to extend species distribution models into the wider ocean at finer scales will enhance the utility and value of marine geomorphology variables for marine biodiversity assessment.

18

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

6.2 Marine geomorphology and geohazards

Early geomorphometric studies of seafloor morphology in the 1960s were limited by the one-dimensionality and the low resolution of the bathymetric data that were available at the time (e.g. Krause and Menard, 1965; Neidell, 1966). In the last three decades, improvements in seafloor surveying technologies have resulted in a renewed interest in employing 5 geomorphometric techniques to study seafloor geomorphology. A list of some works that have utilised geomorphometric techniques in their study of seafloor morphology is provided in Table 2. Similar techniques have also been utilised in the interpretation of sidescan sonar data (e.g. Blondel et al., 1998; Carmichael et al., 1996; Huvenne et al., 2002; Mitchell and Somers, 1989).

Geomorphometric techniques have generally performed well in submarine environments. The use of specific 10 geomorphometric techniques, where features of interest are identified prior to analysis, has involved examining how different morphological parameters change spatially and with each other. They have been amongst the most successful techniques, particularly with regard to the study of submarine mass movements, canyons and volcanoes. In the study of submarine mass movements, the general approach has been the prior identification of the boundaries of the landslides, the measurements of a series of morphometric parameters and their spatial and statistical analyses. This kind of approach has 15 been applied to slope instability offshore Norway (Haflidason et al., 2005; Issler et al., 2005; Micallef et al., 2008), demonstrating scale invariance in terms of statistics and morphology, which has important implications for submarine landslide modelling and hazard assessment. It has also been employed on a finer scale (Casalbore et al., 2011; Rovere et al., 2014) and a broader scale (Hühnerbach et al., 2004; McAdoo et al., 2000; Moernaut and De Batist, 2011) to identify tsunamigenic landslides and to provide interesting insights into failure frequency, preconditioning factors, triggers and 20 controls of submarine mass movements in a wide range of environments, including lakes. In submarine canyons, specific geomorphometric analyses of submarine landslides has shown that landslides can be the most efficient process removing material from canyons and that their influence becomes more significant as the canyon matures (Green and Uken, 2008; Micallef et al., 2012). Geomorphometric investigations of submarine canyon form have generally focused on using morphological data to propose model of canyon erosion by turbidity currents (Mitchell, 2004, 2005; Vachtman et al., 2013). 25 More recently, specific geomorphometric techniques have been used to demonstrate how canyons in passive, progradational margins develop into geometrically self-similar systems that approach steady state and higher drainage efficiency (Micallef et al., 2014b), and how canyons in active margins fail to reach steady-state because of continuous adjustment to perturbations associated with tectonic displacements and base-level change (Micallef et al., 2014a). The geomorphometric study of volcanoes has been useful in determining the key processes constructing and modifying volcano flanks and 30 specifying the conditions that lead to slope instability (Mitchell, 2003; Mitchell et al., 2002; Stretch et al., 2006).

Initially, the techniques of general geomorphometry used in the study of submarine landscapes were less numerous and varied than those used in the study of subaerial landscapes. The majority of studies where geomorphometry was applied to 19

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

the study of submarine landscapes have involved either spectral analyses of the bathymetric data or the statistical analysis of morphometric attributes. More recently, general geomorphometric studies have made wider use of morphometric attributes and their statistical analyses and feature-based quantitative representation, most of which were specifically developed for submarine landscapes. Micallef et al. (2007a), for example, developed a methodology for the quantitative analysis of 5 seafloor data, which was shown to exploit the full potential of these data sets and significantly improve the mapping and characterisation of submarine landslides. This methodology was applied to the submarine mass movements offshore Norway to elucidate the evolution dynamics of a multi-phase submarine landslide (Fig. 8), while emphasising the potential role of gas hydrate dissociation and contourite deposition in controlling the location and extent of submarine slope failure (Micallef et al., 2009), and to improve understanding of the mechanics and triggers of spreading, also while using limit-equilibrium and 10 mechanical modelling (Micallef et al., 2007b). The automated and objective mapping of submarine landscapes is indeed an important application of general geomorphometry, and specific techniques have been developed for the characterisation of pockmarks (Harrison et al., 2011; Gafeira et al., 2012), terraces (Passaro et al., 2011) and canyons (Ismail et al., 2015). Others have used general geomorphometric techniques to classify submarine landscapes (e.g. fjords (Mosalik et al., 2014a; Mosalik et al., 2014b), continental shelf and slope (Elvenes, 2013), and global (Harris et al., 2014)), identify the various 15 styles and scales of deformation across submarine landslides (Mountjoy et al., 2009), and infer the evolution of seamounts (Passaro et al., 2010), mid-ocean ridge scarps (Mitchell et al., 2000) and faults in active continental margins (Kukowski et al., 2008).

6.3 Change detection

A number of studies have described temporal morphological dynamics of the seafloor using acoustic bathymetry (e.g. Duffy 20 & Hughes-Clarke, 2005; Smith et al., 2007). However, assessments of biological change beyond the range of optical sensors have been based primarily on ground sampling methods. Rattray et al. (2013) investigated approaches to quantify temporal change in benthic habitats from a spatially explicit perspective using acoustic techniques. Their methods quantified change in terms of gains and losses in the extent of habitat at a site on the temperate southeast Australian continental shelf; (2) they could distinguish between systematic and random patterns of habitat change; and (3) were able to assess the applicability of 25 supervised acoustic remote sensing methods for broad-scale habitat change assessment. Change detection in temperate bedrock reefs were identified through morphological characterisation by Storlazzi et al. (2013). They delineated the classes using a multivariate classification routine (Dartnell and Gardner, 1999) based on acoustic backscatter and rugosity (surface- planar area ratio).

There have also been several examples in the literature of repeat multibeam surveys being used to detect change, many of 30 which are summarised by Schimel et al. (2015). Analysis is generally focussed on differences in depth values detected and often aided by a visual assessment of the changes in morphology. There are fewer studies that have explicitly used terrain attributes or features in their assessments, but we recognise the potential for gemorphometric techniques to be more widely 20

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

applied in this type of study. For example, Bøe et al. 2015 used geomorphic feature detection (Wood, 1996) to identify crests and ridges in a sandwave field on the continental slope, and assess movement between surveys based on the change in position of these features. We note also that Schimel et al. (2015) incorporate measures of bathymetric uncertainty in their assessment of volume change and also recommend guidelines on thresholds which can help to improve the confidence of 5 such assessments.

6.4 Hydrodynamics

The interaction of bottom currents with seafloor sediments results in a wide range of erosional and depositional morphologies – e.g. scours, furrows, ripples, dunes, lineations, contouritic drifts – the morphology and dimensions of which depend on flow velocity and sediment grain size (Stow et al., 2009). Detecting change in bedform morphology is of great 10 interest to geologists, physical oceanographers and climatologists, and many other with the applied interest in such features. Bedforms determine basic flow patterns of ocean circulation at coarse and fine scales; even small perturbations in seafloor topography can influence major shallow and deep current flows, heat transport and ultimately climate (Gille et al., 2004; Kunze and Llewellyn Smith, 2004; Metzger and Hurlburt, 2001). In turn, bedforms are also excellent archives of current and past bottom flow patterns (Sandwell et al., 2002). Port managers are also interested in bedforms and their evolution, 15 particularly where they constitute a hazard to navigation in coastal waters. Detecting change in bathymetry and its impact on oceanography is therefore important, and local geomorphometric attributes, such as aspect, curvature and rugosity, have been used to develop hydrodynamic models or as proxies for local and regional currents (Lecours et al., 2015a). Seafloor topographic proxies are also fundamental in the predictive mapping of suspension-feeders (Lucieer et al., 2013; Hill et al., 2014), such as cold-water corals (Rengstorf et al., 2012; Tong et al., 2013), because their distribution is inextricably linked to 20 current flow strengths and patterns (Mohn et al., 2014). More recently, understanding the link between seafloor morphometry and currents has been shown to be essential in forecasting the path of floating debris from tsunamis and air disasters and assist in search and rescue operations (Mofield et al., 2004; Normile, 2014; Smith and Marks, 2014).

7 The Future of Marine Geomorphometry

This manuscript has discussed the current practices in marine geomorphometry, from data collection to the applications. 25 Through all aspects of this discussion it is apparent that the use of geomorphometric techniques in the marine realm is relatively nascent, having begun only over a decade ago. The dramatic increase in DBM availability, combined with the increasingly accessible and user-friendly GIS tools, is currently fuelling the amount and diversity of applications of marine geomorphometry. However, this availability can become a double-edge sword. As noted by Dolan and Lucieer (2014), “Although a [DBM] is a model of the seabed surface, it is often not treated as a model but rather is accepted as a true 30 representation of the seabed”. Furthermore, as highlighted in Fig. 1, the end-users of geomorphometric techniques are not 21

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

always aware that they are actually “doing” geomorphometry, but rather think of the steps they are performing as simply using GIS tools for data analysis. As the amount of applications increases, some of the fundamental issues associated with marine geomorphometry are not being addressed quickly or broadly enough. This can increase the risk of unsuspecting end- users misusing data or techniques, and to the misinterpretation of results. For instance, due to a lack of awareness of the 5 impact of artefacts in DBMs and their propagation to terrain attributes, artefacts are often disregarded, or assumed to be obvious. In habitat mapping, the consequences of artefacts are often apparent to geomorphometry-aware users in the final maps (e.g. Zieger et al., 2009; Lucieer et al., 2012), but this can become problematic if the maps are being used in conservation and management decision making if the effect of the artefacts is not appreciated by the end user. We recognise a critical need for a dedicated scientific effort in marine geomorphometry that will address, and raise awareness of the 10 fundamental issues related to marine geomorphometry. This effort does not necessarily have to come solely from the marine science community, indeed it may well benefit from the expertise of many of those scientists already engaged in terrestrial geomorphometry. The main objectives of this effort would be to learn from the lessons of terrestrial geomorphometry, ensure that studies of geomorphometry become more widespread in the marine literature, and respond to the challenges and opportunities for a wider adoption of marine geomorphometry as a key tool in marine sciences, whilst improving and 15 upholding scientific standards.

This section first examines the issues associated with two key growth areas where we recognize particular potential for marine geomorphometry to contribute to our understanding of the seafloor. Following this, we reflect on the current and future trends of the field as a whole, highlighting recommendations to improve practices and set standards for future work.

7.1 The littoral gap: implications for coastal geomorphometry

20 This paper has shown how geomorphometric techniques developed mostly in terrestrial settings can be applied to the marine environment or adapted to enable quantification of the seafloor terrain. The boundary between the land and the sea, however, is not easily mapped or delineated and has always represented a challenge for both marine and terrestrial scientists (Klemas, 2011). Due to the inability of satellite remote sensing to collect data in deep waters and the limitations of acoustic systems to collect data in shallow waters, there is often a gap in terrain data where land meets sea. This littoral gap, sometimes referred 25 to as ‘the white zone’ because the lack of data in this area between available DBM and DEM data appear white on maps, often complicates the study of nearshore environments and can have important implications for applications such as navigation and geohazard assessment. For instance, in their attempt to assess the effectiveness of a marine protected area in a Canadian sub-arctic fjord with habitat maps generated from a combination of terrain attributes and other data, Copeland et al. (2013) were only able to map 32 km2 of the total 82 km2 of the area. They highlighted the laborious nature of shallow water 30 survey (i.e. time and cost-consuming MBES surveys), the need for a continuous coverage because of the large littoral gap, and indicated that interpolation and extrapolation of results in the littoral gap were inappropriate because of the heterogeneous nature of coastal fjord environments (Copeland et al., 2013). 22

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Several examples of investigations of the coastal environment using geomorphometry can be found in the literature (e.g. Purkis et al., 2008; Pittman et al., 2009). However, most coastal studies are limited to small areas either above the water (e.g. dunes or emergent features) or submerged (e.g. coral reefs) (Brock and Purkis, 2009). To our knowledge, there are no geomorphometric applications that span the terrestrial and underwater landscapes in a continuous way over large areas, 5 despite many efforts to collect topo-bathymetric data that span the coastal environment (e.g. National Coastal Mapping Program of the Joint Airborne Lidar Bathymetry Technical Center of Expertise in the United States). For instance, Dunkin et al. (2011) and Dunkin and McCormick (2011) used bathymetric LiDAR datasets to characterise changes in geomorphic and bathymetric features, respectively to assess consequences of storm events, and for monitoring and shore protection purposes. However, these studies did not explicitly use geomorphometric techniques. Several fields could benefit from 10 seamless geomorphometric analysis. For instance, inter-tidal rocky shores are known to shelter a lot of biodiversity (Kostylev et al., 2005) and linking quantitative terrain attributes to measures of biodiversity could improve scientific understanding of ecological patterns and processes in these important areas of the land-sea boundary (e.g. Collin et al., 2012). Observations of underwater and terrestrial landforms have shed light on how glaciers retreated in Atlantic Canada during the last deglaciation (Shaw et al., 2006); the investigation of landforms that overlap both realms could help refine this 15 type of analysis.

At a broader scale, a seamless analysis of terrestrial and marine environments requires the combination of terrestrial DTMs, bathymetric data from acoustic systems, and bathymetric LiDAR or optical remotely sensed data to fill the littoral gap and create a Coastal Terrain Model (CTM) (Hogrefe et al., 2008; Leon et al., 2013). The challenges encountered with merging datasets from different sources makes such an approach still nascent in the general literature (Macon et al., 2008; Quadros et 20 al., 2008; Collin et al., 2012), and very rare, if not fully absent, in the marine geomorphometry literature. Data fusion is the process of acquiring, processing and synergistically combining multi-source datasets both geometrically (i.e. in space) and topologically (i.e. in terms of their attributes or information content) (Usery et al., 1995; Samadzadeghan, 2004; Mohammadi et al., 2011). Despite constant developments in data fusion (Pohl and van Genderen, 1998; Dong et al., 2009; Zhang, 2010), it presents particular challenges for geomorphometry. First, despite improvements in edge matching algorithms, artefacts 25 from merging and surveying can appear when deriving terrain attributes from the fused dataset (Stoker et al., 2009). Data fusion often requires the different datasets to overlap slightly in order to be combined. In theory, the overlapping areas should yield very similar values, within their uncertainty and error ranges. However, important inconsistencies (up to 6.5m) have been reported between depth measurements of the same areas using bathymetric LiDAR and MBES (Quadros et al., 2008; Costa et al., 2009; Chust et al., 2010; Shih et al., 2014). This has implications for geomorphometry since terrain 30 attributes will capture and classify these mismatches as features, especially as the differences usually occur locally (Chust et al. 2010). Also, coastal environments can be very dynamic; artefacts could appear in the DTM if the multi-source data are not collected at the same time and changes occurred between the data collections. Another issue concerns vertical datums (Hogrefe et al., 2008; Erikstad et al., 2013); terrestrial surveys are usually referenced to a local geoid model based on the 23

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

GPS, while underwater surveys (acoustic and LiDAR) are usually referenced to the mean sea level at the time of survey, which is referenced to a local or regional tidal gauge that is itself referenced to a local datum. Calls for a consistent and unified vertical datum have been made but this issue is still unresolved (Hogrefe et al., 2008; Quadros et al., 2008). Finally, data quality and uncertainty may complicate the fusion of the different datasets. For instance, the inability of bathymetric 5 LiDAR systems to collect reliable data in turbid or cloudy waters and in breaking wave conditions, in addition to their difficulties to distinguish the seafloor from the water surface in waters shallower than 30 cm (Quadros et al., 2008) may create a smaller littoral gap called the “dead zone” (Nayegandhi et al., 2009) and prevent proper fusion. Bernstein et al. (2011) recommend a customized survey design to minimise the challenges associated with creating a seamless DTM.

Regular problems of data fusion, for instance related to merging multi-resolution datasets or to software and format 10 compatibility/interoperability, also apply to the development of DTMs for coastal geomorphometry. Terrestrial terrain models may have a format (.dem), while bathymetric LiDAR data can be recorded with a Laser File format (.las) and finally, acoustic data can be save as a Bathymetric Attributed Grid format (.bag); all these file formats have different structure and characteristics. Impediments to the fusion of multisensor data to build seamless elevation and depth surfaces include, but are not limited to, inconsistent spatial and temporal scales, incompatible formats, and differences in 15 levels of reliability, uncertainty and completeness. Despite these impediments, data fusion has been identified as a promising technique for geomorphometry (Bishop et al., 2012). Some authors (e.g. Quadros et al., 2008) argue that the different types of datasets cannot be readily integrated, but the main challenges will likely be addressed with improvements in data fusion techniques and ease of implementation of these techniques for non-expert users (Zhang, 2010) for geomorphometry. Current work includes detection and correction of differences in geoid models, consideration of uncertainties, and improvement in 20 edge matching algorithms (Quadros et al., 2008; Dong et al., 2009; Stoker et al., 2009). Developments in data fusion will likely allow better integration of different data to create seamless coverage for complete geomorphometric analysis and identification of overlapping landforms between the different realms. This will be useful for a wide range of coastal applications, including the investigation of coastal morphodynamics and land-sea exchange modelling, dredging and the identification of hazards due to sea-level rise and severe storms, coastal archaeology, resource management and marine 25 spatial planning, anthropogenic sensitivity and environmental status assessment, and other scientific research.

7.2 Seismic geomorphometry

Seismic geomorphology is a rapidly evolving discipline. It comprises the application of geomorphological principles and analytical techniques to study palaeo-landscapes as imaged by 3D seismic reflection data (Carter, 2003; Posamentier and Kolla, 2003; Posamentier, 2003). More recently, 3D seismic reflection data have also provided a good alternative source of 30 bathymetric data when the latter are absent (e.g. broad scale geomorphic mapping in the MAREANO project – http://www.mareano.no/). The development of seismic geomorphometry is a natural consequence of the advent of high quality and increasingly more affordable 3D seismic reflection datasets, and of increasing computer power, which enables 24

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

the rapid manipulation, visualisation and interpretation of 3D seismic reflection data. The integration of seismic geomorphology with seismic stratigraphy currently represents the state-of-the-art approach to extracting geological information from 3D seismic reflection data to understand large-scale basin evolution. Seismic geomorphological studies have addressed a broad range of geological problems, ranging from sedimentary to igneous geology, from lithology 5 distribution to large-scale tectonic analysis (e.g. Fachmi and Wood, 2003; Miall, 2003; Wood, 2003).

Up to the present, most studies have focused on the qualitative recognition of broad scale features (e.g. Posamentier et al., 1996; Posamentier et al., 2000; Peyton and Boettcher, 2000; Posamentier, 2003; Zeng and Hentz, 2004). Quantitative seismic geomorphology, or seismic geomorphometry, is the most recent development of seismic geomorphology (Carter, 2003; Posamentier, 2003; Posamentier and Kolla, 2003). Seismic geomorphometry has been defined as the “quantitative 10 analysis of the landforms, imaged in 3D seismic data, for the purposes of understanding the history, processes and fill architecture of a basin” (Wood, 2003). Seismic geomorphometry encompasses techniques that use 3D seismic data to investigate the nature and architecture of reservoirs through extraction and analysis of quantitative morphometric information.

Great opportunities exist for applying a more quantitative approach in seismic geomorphology. Seismic geomorphometric 15 techniques provide statistical and mathematical insight into the morphological and dimensional characteristics of geologic systems that are difficult to derive through qualitative investigations of outcrop exposures and 2D seismic reflection data. Seismic geomorphometric studies provide a deep and spatially extensive understanding of how morphology develops through time, providing insight into the historical evolution of a basin and the possibility of developing predictive models. Quantitative relationships derived from seismic geomorphological studies can decrease our uncertainty in predicting the 20 nature and location of reservoirs in deep-water settings by testing cause-and-effect relationships in a variety of settings. Computer-assisted seismic geomorphometry, in particular planform pattern recognition, is a powerful addition to the seismic geomorphological approach. It allows the interpreter to identify geologically significant features in plan view automatically. The ability to exploit the full potential of large seismic data sets is currently hindered by the lack of tools in existing software packages, coupled with the limited knowledge of how morphometrics can be used in the analytical process. It is the 25 development of such tools that should be a main focus for researchers of marine geomorphometry in the near future.

7.3 Current and future trends in marine geomorphometry

Current developments in the marine geomorphometry literature are primarily focussed on the data acquisition end of the workflow. Technology and equipment for surveying the seafloor are improving in quality, accuracy and cost-effectiveness, which will allow an increase in data availability and quality. In coastal environments, ongoing research is focused on 30 improving the extraction of depth information in the littoral gap in order to create seamless DTMs from the seafloor to land. As the pressure on coastal environments increases, such information will become crucial for many applications. From an

25

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

ecosystem point of view, coastal environments are also very rich in biodiversity. Studies of the topographic structure that can be identified from CTMs using geomorphometric techniques are likely to facilitate a better scientific understanding of these ecosystems. In the deep-sea, extensive use of AUV and ROV-based MBES and other technologies means we are now able to collect high-resolution bathymetric data of environments never explored before at such level of detail. The knowledge that 5 has been gained from using these data in combination with different techniques, including geomorphometry, has revolutionized scientific understanding of many marine environments. It was initially thought that the deep-sea was mostly flat, muddy and lifeless, but the last twenty years of research have proven otherwise. Nevertheless, exploration is far from complete; there are still wide gaps in the scientific knowledge of deep-sea patterns, processes and ecosystems. High quality bathymetric data is fundamental to the success of revealing this knowledge and its limited availability is currently a barrier to 10 effective protection and management of vulnerable species (Vierod et al., 2014; Ross et al., 2015).

As the marine geomorphometry community moves forward, it will rapidly need to start addressing issues other than those associated with data acquisition. The availability of tools that streamline the workflow from data collection to analysis will be key in making a more complete science of marine geomorphometry accessible to marine scientists with a wide range of background and experience. Repositories of comprehensive and freely available datasets and tools, such as Digital Coast that 15 provide free coastal and marine bathymetric data and analytical tools (NOAA, 2016), are the way forward to improve accessibility to the wider scientific community, and this may well mean that bathymetric data gain the attention of those currently engaged in developing geomorphometric methods for terrestrial data. We also acknowledge that easily accessible GIS tools and readily available data can also bring hidden dangers from non-critical use by users with limited appreciation of data collection and processing methods which to the expert clearly reflect the limitations in the utility of particular 20 bathymetric datasets. To prevent this danger of inappropriate use, tools and datasets need to be accompanied by complete metadata that include information on data provenance, survey, scale, error and uncertainty quantification, and any other information relevant to further use of the tools and datasets. Metadata are crucial to create a “quality-aware” community (Devillers et al., 2007; Lecours and Devillers, 2015; Lecours et al., 2015b). The use of the CUBE algorithm to create BASE surfaces is one way to carry over a measure of quantified uncertainty of the data, but such information is not readily 25 available for the majority of publicly available datasets. This type of information needs to become more accessible to marine scientists with a broad range of scientific backgrounds.

As indicated in Fig. 1, it is becoming critical to raise awareness of geomorphometry in the marine science community. Methods from specific geomorphometry demonstrate a lot of potential for marine geomorphometry and should be used more extensively. At the same time, it is opportune to improve practices by setting standards and protocols for the application of 30 marine geomorphometry. Methods and interpretations need to be standardised, particularly in view of issues specific to the marine environment, or where data and analyses behave differently underwater than on land. Amongst these, the influence of scale and data resolution on the results, and the consideration of spatial uncertainty should be prioritised. Since geomorphometric analyses are more and more performed within GIS environments, devising a GIS-based standard 26

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

methodology and symbology for marine geomorphological mapping using geomorphometry would be a very useful goal for the marine geomorphometry community. Ultimately, the type of standards and protocols a marine geomorphometry community could develop should encourage wider applications of bathymetric data and allow marine scientists optimise the use of their expensive datasets.

5 8 Conclusions

Unlike terrestrial applications of geomorphometry, which can be traced back to the nineteenth century, the use of geomorphometry in the marine realm is still in its infancy. Ever since the first coarse-scale DBMs were generated, marine geomorphometry has helped improve scientific understanding of the oceans, from the relatively thin border where land meets sea to the deepest waters. This paper is timely because it reviews the state of the art in the field and discusses 10 standards for the applications of marine geomorphometry. We have reviewed the different methods to sample the depth of the seafloor, the interpolation methods and issues of spatial scale associated with the generation of a DBM, the different errors and artefacts that are characteristics of DBMs and different from DEMs, and discussed how general and specific geomorphometry are applied underwater. We also introduced four common and two cutting edge applications of marine geomorphometry and discussed the future trends of the field. Based on this review, we provide the following 15 recommendations that will lead to more efficient practices in marine geomorphometry: (1) errors and spatial uncertainty should be quantified so that they are able to be considered in the geomorphometric analyses and in the interpretation of results; (2) metadata should consistently be associated with datasets to explicitly indicate data provenance, quality (i.e. quantification of uncertainty), and the spatial scale at which the dataset was intended to be used; (3) data, metadata and tools should be made available for a wider applications of bathymetric data; (4) standardisation methods and interpretations for 20 each field of application should be documented, particularly in view of the influence of scale and data resolution on the results; (5) a GIS-based standard symbology for marine geomorphological mapping based on geomorphometry should be devised.

By raising awareness, we hope that both marine scientists and geomorphometry practitioners will work together in addressing the fundamental issues of marine geomorphometry, whilst upholding the standards of the discipline of marine 25 . Building a dedicated effort in marine geomorphometry that can draw from lessons learned in terrestrial geomorphometry and encourage marine applications will ensure that geomorphometry as a science is able to fulfil its potential.

27

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Acknowledgements

V. Lecours thanks Emma LeClerc for her valuable comments on sections of this manuscript, Dr. Rodolphe Devillers for the insightful discussions about marine geomorphometry, and the Natural Sciences and Engineering Research Council (NSERC) of Canada for providing funding. A. Micallef is funded by a Marie Curie Career Integration Grant PCIG13-GA-2013- 5 618149.

References

Adams, E. W., and Schlager, W.: Basic types of submarine slope curvature. J. Sediment. Res., 70, 814-828, 2000.

Anderson, M. G., and Ferree, C. E.: Conserving the stage: climate change and the geophysical underpinnings of species diversity. PLOS One, 5, e11554, 2010.

10 Arifin, R. R., and Kennedy, A. B.: The evolution of large scale crescentic bars on the northern Gulf of Mexico coast. Mar. Geol., 285, 46-58, 2011.

Becker, J. J., Sandwell, D. T., Smith, W. H. F., Braud, J., Binder, B., Depner, J., Fabre, D., Factor, J., Ingalls, S., Kim, S.-H., Ladner, R., Marks, K., Nelson, S., Pharaoh, A., Trimmer, R., Von Rosenberg, J., Wallace, G., and Weatherall, P.: Global bathymetry and elevation data at 30 arc seconds resolution: SRTM30_PLUS. Mar. Geod., 32, 355–371, 2009.

15 Benz, U. C., Hofmann, P., Willhauck, G., Lingenfelder, I., and Heynen, M.: Multi-resolution, object-oriented fuzzy analysis of remote sensing data for GIS-ready information. ISPRS J. Photogramm., 58, 239-258, 2004.

Berkson, J. M., and Matthews, J. E.: Statistical properties of seafloor roughness, in: Acoustics in the Sea-bed, Bath University Press, Bath, United Kingdom, 215-223, 1983.

Bernstein, D. J., Sumners, B. J., Freeman, C. W., and Mitasova, H.: Modern techniques for improved topo/bathy elevation 20 modelling of tidal inlets. U.S. Hydro 2011, 1-11, 2011.

Bishop, M. P., James, L. A., Shroder Jr., J. F., and Walsh, S. J.: Geospatial technologies and digital geomorphological mapping: concepts, issues and research. Geomorphology, 137, 5-26, 2012.

Blaschke, T., Hay, G. J., Kelly, M., Lang, S., Hofmann, P., Addink, E., Queiroz Feitosa, R., van der Meer, F., van der Werff, H., van Coillie, F., and Tiede, D.: Geographic object-based image analysis – towards a new paradigm. ISPRS J. 25 Photogramm., 87, 180-191, 2014.

Blaszczynski, J. S.: Landform characterization with geographic information systems. Photogramm. Eng. Rem. S., 63, 183- 191, 1997.

28

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Blondel, P., Parson, L. M., and Robigou, V.: TexAn: Textural analysis of sidescan sonar imagery and generic seafloor characterisation. Oceans-IEEE, 419-423, 1998.

Booth, J. S., and O'Leary, D. W.: A statistical overview of mass movement characteristics on the North American Atlantic Outer Continental Margin. Mar. Geotechnol., 10, 1-18, 1991.

5 Brock, J. C., and Purkis, S. J.: The emerging role of LiDAR remote sensing in coastal research and resource management. J. Coastal Res., 53, 1-5, 2009.

Brown, C. J., Sameoto, J. A., and Smith, S. J.: Multiple methods, maps, and management applications: purpose made seafloor maps in support of ocean management. J. Sea Res., 72, 1-13, 2012.

Brown, C. J., Smith, J. A., Lawton, P., and Anderson, J. T.: Benthic habitat mapping: a review of progress towards improved 10 understanding of the spatial ecology of the seafloor using acoustic techniques. Estuar. Coast. Shelf S., 92, 502-520, 2011.

Bue, B. D., and Stepinski, T. F.: Automated classification of landforms on Mars. Comput. Geosci., 32, 604-614, 2006.

Bøe, R., Skarðhamar, J., Rise, L., Dolan, M. F., Bellec, V. K., Winsborrow, M., Skagseth, Ø., Knies, J., King, E. L., Walderhaug, O., and Chand, S.: Sandwaves and sand transport on the Barents Sea continental slope offshore northern Norway. Mar. Petrol. Geol., 60, 34-53, 2015.

15 Buscarnera, G., and Di Prisco, C.: Soil stability and flow slides in unsaturated shallow slopes: can saturation events trigger liquefaction processes? Geotechnique, 63, 801-817, 2013.

Calder, B.: Automatic statistical processing of multibeam echosounder data. Int. Hydrogr. Rev., 4, 53-68, 2003.

Calder, B., Mayer, L. A.: Automatic processing of high-rate, high-density multibeam echosounder data. Geochem. Geophy. Geosy., 4, 1048-1070, 2003.

20 Calmant, S., and Baudry, N.: Modelling bathymetry by inverting satellite altimetry data: a review. Mar. Geophys, Res., 18, 123-134, 1996.

Calvert, J., Strong, J. A., Service, M., McGonigle, C., and Quinn, R.: An evaluation of supervised and unsupervised classification techniques for marine benthic habitat mapping using multibeam echosounder data. ICES J. Mar. Sci., 72, 1498- 1513, 2014.

25 Cameron, M. J., Lucieer, V., Barrett, N. S., Johnson, C. R., and Edgar, G. J.: Understanding community-habitat associations of temperate reef fishes using fine-resolution bathymetric measures of physical structure. Mar. Ecol. Prog. Ser., 506, 213- 229, 2014.

Carbonneau, P. E., Lane, S. N., and Bergeron, N.: Feature based image processing methods applied to bathymetric measurements from airborne remote sensing in fluvial environments. Earth Surf. Proc. Land., 31, 1413-1423, 2006. 29

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Carmichael, D. R., Linnett, L. M., Clarke, S. J., and Calder, B. R.: Seabed classification through multifractal analysis of sidescan sonar imagery. IEE P-Radar Son. Nav., 143, 140-148, 1996.

Carr, M. H., Greeley, G., Blasius, K. R., Guest, J. E., and Murray, J. B.: Some Martian features as viewed from the Viking orbiter. J. Geophys, Res., 82, 3985-4015, 1977.

5 Carrara, A., Bitelli, G., and Carla, R.: Comparison of techniques for generating digital terrain models from contour lines. Int. J. Geogr. Inf. Sci., 11, 451-473, 1997.

Carron, M. J., Vogt, P. R., and Jung, W.-Y.: A proposed international long-term project to systematically map the world’s ocean floors from beach to trench: GOMaP (Global Ocean Mapping Program), Int. Hydrogr. Rev., 2, 49–55, 2001.

Carter, D. C.: Channel and sandstone body geometry from 3D seismic and well control in Widuri Field, Offshore SE 10 Sumatra, Indonesia. AAPG Bull., 87, 909-934, 2003.

Casalbore, D., Romagnoli, C., Bosman, A., and Chiocci, F.: Potential tsunamigenic landslides at Stromboli Volcano (Italy): Insight from marine DEM analysis. Geomorphology, 126, 42-50, 2011.

Chakraborty, B., Schenke, H. W., Kodagali, V., and Hagen, R.: Analysis of multibeam-Hydrosweep echo peaks for seabed characterisation. Geo-Mar. Lett., 20, 174-181, 2001.

15 Chorley, R. J., Malm, D. E. C., and Pogorzelski, H. A.: A new standard for measuring drainage basin shape. Am. J. Sci., 255, 138-141, 1957.

Chust, G., Grande, M., Galparsoro, I., Uriarte, A., and Borja, Á.: Capabilities of the bathymetric Hawk Eye LiDAR for coastal habitat mapping: a case study within a Basque estuary. Estuar. Coast. Shelf S., 89, 200-213, 2010.

Coggan, R. A., and Diesing, M.: Rock ridges in the central English Channel, in: Seafloor geomorphology as benthic habitat: 20 GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 471-480, 2012.

Collier, J. S., and Humber, S. R.: Fringing reefs of the Seychelles inner granitic islands, western Indian Ocean, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 340-347, 2012.

Collin, A., Long, B., and Archambault, P.: Merging land-marine realms: spatial patterns of seamless coastal habitats using a 25 multispectral LiDAR. Remote Sens. Environ., 123, 390-399, 2012.

Copeland, A., Edinger, E., Devillers, R., Bell, T., LeBLanc, P., and Wroblewski, J.: Marine habitat mapping in support of Marine Protected Area management in a subarctic fjord: Gilbert Bay, Labrador, Canada. J. Coast. Conserv., 17, 225-237, 2013.

30

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Costa, B. M., and Battista, T. A.: The semi-automated classification of acoustic imagery for characterizing coral reef ecosystems. Int. J. Remote Sens., 34, 6389-6422, 2013.

Costa, B. M., Battista, T. A., and Pittman, S. J.: Comparative evaluation of airborne LiDAR and ship-based multibeam SoNAR bathymetry and intensity for mapping coral reef ecosystems. Remote Sens. Environ., 113, 1082-1100, 2009.

5 Cushman-Roisin, B., and Beckers, J.-M.: Introduction to geophysical fluid dynamics, 2nd edition, Academic Press, 875 p., 2011.

Dartnell, P., and Gardner, J. V.: Predicting seafloor facies from multibeam bathymetry and backscatter data. Photogramm. Eng. Rem. S., 70, 1081-1091, 2004.

Davies, A. J., Wisshak, M., Orr, J. C., and Roberts, J. M.: Predicting suitable habitat for the cold-water coral Lophelia 10 pertusa (Scleractinia). Deep-Sea Res. Pt I, 55, 1048-1062, 2008.

Dekker, A. G., Phinn, S. R., Anstee, J., Bissett, P., Brando, V. E., Casey, B., Fearns, P., Hedley, J., Klonowski, W., Lee, Z. P., Lynch, M., Lyons, M., Mobley, C., and Roelfsema, C.: Intercomparison of shallow water bathymetry, hydro-optics, and benthos mapping techniques in Australian and Caribbean coastal environments. Limnol. Oceanogr.-Meth., 9, 396-425, 2011.

De Moustier, C., and Matsumoto, H.: Seafloor acoustic remote sensing with multibeam echo-sounders and bathymetric 15 sidescan sonar system. Mar. Geophys. Res., 15, 27-42, 1993.

Devillers, R., Bédard, Y., Gervais, M., Jeansoulin, R., Pinet, F., Schneider, M., Bejaoui, L., Lévesque, M.-A., Salehi, M., and Zargar, A. : How to improve geospatial data usability: from metadata to quality-aware GIS community, in: Spatial data usability, an AGILE pre-conference workshop, Proceedings of the 10th AGILE International Conference on Geographic Information Science, Aalborg, AGILE (Association of Geographic Information Laboratories in Europe), 1-8, 2007.

20 Diesing, M., Green, S. L., Stephens, D., Lark, R. M., Stewart, H. A., and Dove, D.: Mapping seabed sediments: comparison of manual, geostatistical, object-based image analysis and machine learning approaches. Cont. Shelf Res., 84, 107-119, 2014.

Dimri, A. P., Yasunari, T., Wiltshire, A., Kumar, P., Mathison, C., Ridley, J., and Jacob, D.: Application of regional climate models to the Indian winter monsoon over the western Himalayas. Sci. Total Environ., 468-469, S36-S47, 2013.

25 Dixon, T. H., Naraghi, M., McNutt, M. K., and Smith, S. M.: Bathymetric prediction from SEASAT altimeter data. J. Geophys, Res., 88, 1563-1571, 1983.

Dolan, M. F., Grehan, A. J., Guinan, J. C., and Brown, C.: Modelling the local distribution of cold-water corals in relation to bathymetric variables: Adding spatial context to deep-sea video data. Deep Sea Res. Pt I, 55, 1564-1579, 2008.

31

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Dolan, M. F. J., Thorsnes, T., Leth, J., Al-Hamdani, Z., Guinan, J., and Van Lancker, V.: Terrain characterization from bathymetry data at various resolutions in European waters – experiences and recommendations. NGU Report 2012.045, Geological Survey of Norway, Trondheim, 2012.

Dolan, M. F. J., and Lucieer, V. L.: Variation and uncertainty in bathymetric slope calculations using geographic information 5 systems. Mar. Geod., 37, 187-219, 2014.

Dong, J., Zhuang, D., Huang Y., and Fu, J.: Advance in multi-sensor data fusion: algorithms and applications. Sensors, 9, 7771-7784, 2009.

Douglas, B. C., McAdoo, D. C., and Cheney, R. E.: Oceanographic and geophysical applications of satellite altimetry. Rev. Geophys., 25, 875-880, 1987.

10 Drăguţ, L., and Blaschke, T.: Automated classification of landform elements using object-based image analysis. Geomorphology, 81, 330-344, 2006.

Drăguţ, L., Eisank, C., and Strasser, T.: Local variance for multi-scale analysis in geomorphometry. Geomorphology, 130, 162-172, 2011.

Duffy, G. P., and Hughes-Clarke, J. E.: Application of spatial cross correlation to detection of migration of submarine sand 15 dunes. J. Geophys. Res. F-Ea. Surf., 110, 1-11, 2005.

Dunkin, L. M., and McCormick, J.: Monitoring shore protection projects along the Florida panhandle using three- dimensional spatial data. Coastal Engineering Practice – Proceedings of the 2011 Conference on Coastal Engineering Practice, 742-752, 2011.

Dunkin, L. M., Wozencraft, J. M., and McCormick, J.: Changes to geomorphic and bathymetric features due to extreme 20 storm events near East Pass inlet, Florida. Solutions to Coastal Disasters 2011, 176-187, 2011.

Dunn, D. C., and Halpin, P. N.: Rugosity-based regional modeling of hard-bottom habitat. Mar. Ecol. Prog. Ser, 377, 1-11, 2009.

Du Preez, C.: A new arc–chord ratio (ACR) rugosity index for quantifying three-dimensional landscape structural complexity. Landscape Ecol., 30, 181-192, 2015.

25 Eisank, C., Smith, M., and Hillier, J.: Assessment of multiresolution segmentation for delimiting drumlins in digital elevation models. Geomorphology, 214, 452-464, 2014.

Elvenes, S.: Landscape mapping in MAREANO. NGU report 2013.035, Geological Survey of Norway, Trondheim, 2013.

Elvenes, S., Dolan, M. F., Buhl-Mortensen, P., and Bellec, V. K.: An evaluation of compiled single-beam bathymetry data as a basis for regional sediment and biotope mapping. ICES J. Mar. Sci., 71, 867-881, 2014. 32

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

EMODnet: European Marine Observation and Data Network (EMODnet) bathymetry portal, http://www.emodnet- hydrography.eu/, 2015.

Erikstad, L., Bakkestuen, V., Bekkby, T., and Halvorsen, R.: Impact of scale and quality of digital terrain models on predictability of seabed terrain types. Mar. Geod., 36, 2-21, 2013.

5 Eugenio, F., Marcello, J., and Martin, J.: High-resolution maps of bathymetry and benthic habitats in shallow-water environments using multispectral remote sensing imagery. IEEE T. Geosci. Remote, 53, 3539-3549, 2015.

Evans, I. S.: General geomorphometry, derivatives of altitude, and descriptive statistics, in Spatial Analysis in Geomorphology, Methuen, London, UK, 17–90, 1972.

Evans, I. S: An integrated system of terrain analysis and slope mapping. Z Geomorphol. Supp., 36, 274-295, 1980.

10 Evans, I. S., and Minár, J.: A classification of geomorphometric variables, in Geomorphometry 2011, Redlands, United States, 105-108, 2011.

Ezhova, E., Dorokhov, D., Sivkov, V., Zhamoida, V., Ryabchuk, D., and Kocheshkova, O.: Benthic habitats and benthic communities in Southeastern Baltic Sea, Russian sector, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 613-621, 2012.

15 Fachmi, M., and Wood, L.: Quantitative seismic geomorphology of Belanak and Gabus fields, West Natuna Basin, Indonesia. AAPG Ann. Meet., 51–52, 2003.

Fernandes, M. J., Bastos, L., and Catalão, J.: The role of multi-mission ERS altimetry in the determination of the marine geoid in the Azores. Mar. Geod., 23, 1-16, 2000.

Feurer, D., Bailly, J.-S., Puech, C., Le Coarer, Y., and Viau, A. A.: Very-high-resolution mapping of river-immersed 20 topography by remote sensing. Prog. Phys. Geog., 32, 403-419, 2008.

Fisher, P. F., and Tate, N. J.: Causes and consequences of error in digital elevation models. Prog. Phys. Geog., 30, 467-489, 2006.

Florinsky, I. V.: Accuracy of local topographic variables derived from digital elevation models. Int. J. Geogr. Inf. Sci., 12, 47-61, 1998.

25 Florinsky, I. V.: Digital terrain analysis in soil science and geology. Elsevier/Academic Press, The Netherlands, 379 p., 2012.

Fonstad, M. A., and Marcus, W. A.: Remote sensing of stream depths with hydraulically assisted bathymetry (HAB) models. Geomorphology, 72, 320-339, 2005.

33

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Ford, P. G., and Pettengill, G. H.: Venus topography and kilometre-scale slopes. J. Geophys, Res., 97, 13103-13114, 1992.

Fosså, J. H., Lindberg, B., Christensen, O., Lundälv, T., Svellingen, I., Mortensen, P. B.., and Alvsvag, J.: Mapping of Lophelia reefs in Norway: experiences and survey methods, in: Cold-water corals and ecosytems, Springer-Verlag, Berlin, Germany, 359-391, 2005.

5 Foster, G., Walker, B. K., and Riegl, B. M.: Interpretation of single-beam acoustic backscatter using LiDAR-derived topographic complexity and benthic habitat classifications in a coral reef environment. J. Coastal Res., 53, 16-26, 2009.

Fox, C. G.: Objective classification of oceanic ridge-crest terrains using two-dimensional spectral models of bathymetry: Application to the Juan de Fuca Ridge. Mar. Geophys. Res., 18, 707-728, 1996.

Fox, C. G., and Hayes, D. E.: Quantitative methods for analyzing the roughness of the seafloor. Rev. Geophys., 23, 1-48, 10 1985.

Friedman, A., Pizarro, O., Williams, S. B., and Johnson-Roberson, M.: Multi-scale measures of rugosity, slope and aspect from benthic stereo image reconstructions. PLOS One, 7, e50440, 2012.

Fu, L.-L.: Recent progress in the application of satellite altimetry to observing the mesoscale variability and the general circulation of the oceans. Rev. Geophys. Space Phys., 21, 1657-1666, 1983.

15 Gafeira, J., Long, D., Diaz-Doce, D.: Semi-automated characterisation of seabed pockmarks in the central North Sea. Near. Surf. Geophys., 10, 303-14, 2012.

Galparsoro, I., Borja, Á., Bald, J., Liria, P., and Chust, G.: Predicting suitable habitat for the European lobster (Homarus gammarus), on the Basque continental shelf (Bay of Biscay), using Ecological-Niche Factor Analysis. Ecol. Model., 220, 556-567, 2009.

20 GEBCO: General bathymetric chart of the oceans 30 arc-second global grid of elevations 2014. www.gebco.net, 2014.

Gee, M. J. R., Watts, A. B., Masson, D. G., and Mitchell, N. C.: Landslides and the evolution of El Hierro in the Canary Islands. Mar. Geol., 177, 271-293, 2001.

Gessler, P., Pike, R., MacMillan, R. A., Hengl, T., and Reuter, H. I.: The future of geomorphometry, in: Geomorphometry: concepts, software, applications, Elsevier, Amsterdam, The Netherlands, 637-652, 2009.

25 Gilbert, L. E., and Malinverno, A.: A characterization of the spectral density of residual ocean floor topography. Geophys. Res. Lett., 15, 1401-1404, 1988.

Gille, S. T., Metzger, E. J., and Tokmakian, R.: Seafloor topography and ocean circulation. Oceanography, 17, 47-54, 2004.

34

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Gjerde, K. M.: Ecosystems and biodiversity in deep waters and high seas. UNEP Regional Seas Reports and Studies 178, UNEP/IUCN, Gland, Switzerland, 2006.

Goff, J. A., and Jordan, T. H.: Stochastic modelling of seafloor morphology: Inversion of Sea Beam data for second-order statistics. J. Geophys. Res., 93, 13589-13608, 1988.

5 Goff, J. A., and Tucholke, B. E.: Multiscale spectral analysis of bathymetry on the flank of the Mid-Atlantic Ridge: Modification of the seafloor by mass wasting and sedimentation. J. Geophys. Res., 102, 15447-15462, 1997.

Goff, J. A., Smith, W. H. F., and Marks, K. M.: The contributions of abyssal hill morphology and noise to altimetric gravity fabric. Oceanography, 17, 24-37, 2004.

Graham, D., Chen, A., and Ermisch, D.: NOAA shoreline data explorer application. Coastal GeoTools 2015, March 30-April 10 2, Charleston, North Carolina, 2015.

Green, A., and Uken, R.: Submarine landsliding and canyon evolution on the northern KwaZuluNatal continental shelf, South Africa. Mar. Geol., 254, 152-170, 2008.

Greene, H. G., Yoklavich, M. M., Starr, R. M., O’Connell, V. M., Wakefield, W. W., Sullivan, D. E., McRea Jr., J. E., and Cailliet, G. M.: A classification scheme for deep seafloor habitats. Oceanol. Acta, 22, 663-678, 1999.

15 Grohmann, C. H.: Morphometric analysis in geographic information systems: applications of free software GRASS and R. Comput. Geosci., 30, 1055-1067, 2004.

Grohmann, C. H.: Effects of spatial resolution on slope and aspect derivation for regional-scale analysis. Comput. Geosci., 77, 111-117, 2015.

Guenther, G. C., Lillycrop, W. J., and Banic, J. R.: Future advancements in airborne hydrography. Int. Hydrogr. Rev., 3, 67- 20 90, 2002.

Guinan, J., Brown, C., Dolan, M. F. J., and Grehan, A. J.: Ecological niche modelling of the distribution of cold-water coral habitat using underwater remote sensing data. Ecol. Informatics, 4, 83-92, 2009.

Haflidason, H., Lien, R., Sejrup, H.P., Forsberg, C.F., and Bryn, P.: The dating and morphometry of the Storegga Slide. Mar. Petrol. Geol., 22, 123-136, 2005.

25 Hamylton, S. M., Hedley, J. D., and Beaman, R. J.: Derivation of high-resolution bathymetry from multispectral satellite imagery: a comparison of empirical and optimisation methods through geographical error analysis. Remote Sens., 7, 16257- 16273, 2015.

Harris, P. T., and Baker, E. K.: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, The Netherlands, 900 p., 2012a. 35

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Harris, P. T., and Baker, E. K.: Synthesis and lessons learned, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 871-887, 2012b.

Harris, P. T., and Baker, E. K.: Why map benthic habitats?, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 3-22, 2012c.

5 Harris, P. T., Macmillan-Lawler, M., Rupp, J., and Baker, E. K.: Geomorphology of the oceans. Mar. Geol., 352, 4-24, 2014.

Harrison, J. W., Ferrie, F. P., Hefford, S. W., Samson, C., Kusevic, K., Mrstik, P., and Iles, P. J. W.: Finding anomalies in high-density LiDAR point clouds. Geomatica, 63, 397-405, 2009.

Harrison, R., Bellec, V. K., Mann, D., and Wang, W.: A new approach to the automated mapping of pockmarks in multi- beam bathymetry. IEEE Image Proc., 18, 2777-2780, 2011.

10 Hasan, R. C., Ierodiaconou, D., and Laurenson, L.: Combining angular response classification and backscatter imagery segmentation for benthic biological habitat mapping. Estuar. Coast. Shelf S., 97, 1-9, 2012.

Hedley, J. D., Harborne, A. R., and Mumby, P. J.: Simple and robust removal of sun glint for mapping shallow-water benthos. Int. J. Remote Sens., 26, 2107-2112, 2005.

Hengl, T., and Evans, I. S.: Mathematical and digital models of the land surface, in: Geomorphometry: concepts, software, 15 applications, Elsevier, Amsterdam, The Netherlands, 31-63, 2009.

Hengl, T., and MacMillan, R. A.: Geomorphometry – a key to landscape mapping and modelling, in: Geomorphometry: concepts, software, applications, Elsevier, Amsterdam, The Netherlands, 433-460, 2009.

Herzfeld, U. C.: Geostatistical methods for evaluation of seabeam bathymetric surveys: Case studies of Wegener Canyon, Antarctica. Mar. Geol., 88, 83-95, 1989.

20 Herzfeld, U. C., and Higginson, C. A.: Automated geostatistical seafloor classification - principles, parameters, feature vectors and discrimination criteria. Comput. Geosci., 22, 35-52, 1996.

Heyman, W. D., and Kobara, S.: Geomorphology of reef fish spawning aggregrations in Belize and the Cayman Islands (Caribbean), in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 387-396, 2012.

25 Hill, N. A., Lucieer, V., Barrett, N. S., Anderson, T. J., and Williams, S. B.: Filling the gaps: predicting the distribution of temperate reef biota using high resolution biological and acoustic data. Estuar. Coast. Shelf S., 147, 137-147, 2014.

Hogrefe, K. R., Wright, D. J., and Hochberg, E. J.: Derivation and integration of shallow-water bathymetry: implications for coastal terrain modelling and subsequent analyses. Mar. Geod., 31, 299-317, 2008.

36

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Holman, R., and Haller, M. C.: Remote sensing of the nearshore. Annu. Rev. Marine Sci., 5, 95-113, 2013.

Horn, B. K. P.: Hill shading and the reflectance map. P. IEEE, 69, 14-47, 1981.

Hughes Clarke, J. E.: Dynamic motion residuals in swath sonar data: ironing out the creases. Int. Hydrogr. Rev., 4, 6-23, 2003a.

5 Hughes Clarke, J. E.: A reassessment of vessel coordinate systems: what is it that we are really aligning? US Hydrographic Conference 2003, Biloxi, MS, 1-12, 2003b.

Hughes Clarke, J. E., Mayer, L. A., and Wells, D. E.: Shallow-water imaging multibeam sonars: a new tool for investigating seafloor processes in the coastal zone and on the continental shelf. Mar. Geophys. Res., 18, 607-629, 1996.

Hühnerbach, V., Masson, D. G., and Partners of the COSTA-Project: Landslides in the North Atlantic and its adjacent seas: 10 an analysis of their morphology, setting and behaviour. Mar. Geol., 213, 343-362, 2004.

Huvenne, V. A. I., Beyer, A., de Haas, H., Dekindt, K., Henriet, J.-P., Kozachenko, M., Olu-Le Roy, K., Wheeler, A. J., and the TOBI/Pelagia 197 and CARACOLE cruise participants: The seabed appearance of different coral bank provinces in the Porcupine Seabight, NE Atlantic: results from sidescan sonar and ROV seabed mapping, in: Cold-water corals and ecosytems, Springer-Verlag, Berlin, Germany, 535-569, 2005.

15 Huvenne, V. A. I., Blondel, P., and Henriet, J. P.: Textural analyses of sidescan sonar imagery from two mound provinces in the Porcupine Seabight. Mar. Geol., 189, 323-341, 2002.

IHO, International Hydrographic Organisation: IHO Standards for Hydrographic Surveys. Special publication No. 44. 5th Edition, February 2008. International Hydrographic Bureau, Monaco, 2008.

Iordanishvili, I. K.: The peculiarities of hydrological processes in mountain reservoirs. Water Resour., 27, 617-624, 2000.

20 Irish, J. L., and Lillycrop, W. J.: Scanning laser mapping of the coastal zone: the SHOALS system. ISPRS J. Photogramm., 54, 123-129, 1999.

Ismail, K., Huvenne, V. A. I., and Masson, D. G.: Objective automated classification technique for marine landscape mapping in submarine canyons. Mar. Geol., 362, 17-32, 2015.

Issler, D., De Blasio, F. V., Elverhøi, A., Bryn, P., and Lien, R.: Scaling behaviour of clay-rich submarine debris flows. Mar. 25 Petrol. Geol., 22, 187-194, 2005.

Jakobsson, M., Mayer, L., Coakley, B., Dowdeswell, J. A., Forbes, S., Fridman, B., Hodnesdal, H., Noormets, R., Pederson, R., Rebesco, M., Schenke, H. W., Zarayskaya, Y., Accettella, D., Armstrong, A., Anderson, R. M., Bienhoff, P., Camerlenghi, A., Church, I., Edwards, M., Gardner, J. V., Hall, J. K., Hell, B., Hestvik, O., Kristoffersen, Y., Marcussen, C.,

37

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Mohammad, R., Mosher, D., Nghiem, S. V., Pedrosa, M. T., Travaglini, P. G., and Weatherall, P.: The International Bathymetric Chart of the Arctic Ocean (IBCAO) version 3.0. Geophys. Res. Lett., 39, L12609, 2012.

Jalali, M. A., Ierodiaconou, D., Monk, J., Gorfine, H., and Rattray, A.: Predictive mapping of abalone fishing grounds using remotely-sensed LiDAR and commercial catch data. Fish. Res., 169, 26-36, 2015.

5 James, J. W. C., Pearce, B., Coggan, R. A., and Morando, A.: Open shelf valley system, northern palaeovalley, English Channel, UK, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 587-596, 2012.

Jayne, S. R., St. Laurent, L. C., and Gille, S. T.: Connections between ocean bottom topography and Earth’s climate. Oceanography, 17, 65-74, 2004.

10 Jenness, J. S.: Calculating landscape surface area from digital elevation models. Wildlife Soc. B., 32, 829-839, 2004.

Jensen, J. R.: Introductory digital image processing, a remote sensing perspective, 3rd edition, Pearson Prentice Hall, Upper Saddle River, United States, 526 p., 2005.

Jiang, M., Stewart, W., and Marra, M.: Segmentation of seafloor sidescan imagery using Markov random fields and neural networks. Oceans-IEEE, 456-461, 1993.

15 Johnson-Roberson, M., Pizarro, O., Williams, S. B., and Mahon, I.: Generation and visualization of large-scale three- dimensional reconstructions from underwater robotic surveys. J. Field Robot., 27, 21-51, 2010.

Jones, K. H.: A comparison of algorithms used to compute hill slope as a property of the DEM. Comput. Geosci., 24, 315- 323, 1998.

Kay, S., Hedley, J., and Lavender, S.: Sun glint correction of high and low spatial resolution images of aquatic scenes: a 20 review of methods for visible and near-infrared wavelengths. Remote Sens., 1, 697-730, 2009.

Keating, B., Cherkis, N. Z., Fell, P. W., Handschmacher, D., Hey, R. N., Lazarewicz, A., Naar, D. F., Perry, R. K., Sandwell, D., Schwank, D. C., Vogt, P. R., and Zondek, B.: Field tests of SEASAT bathymetric detections. Mar. Geophys. Res., 7, 69- 71, 1984.

Kennedy, D. M., Ierodiaconou, D., and Schimel, A.: Granitic coastal geomorphology: applying integrated terrestrial and 25 bathymetric LiDAR with multibeam sonar to examine coastal landscape evolution. Earth Surf. Proc. Land., 39, 1663-1674, 2014.

Kenny, A. J., Cato, I., Desprez, G., Fader, T., Schüttenhelm, R. T. E., and Side, J.: An overview of seabed-mapping technologies in the context of marine habitat classification, ICES J. Mar. Sci., 60, 411–418, 2003.

Klemas, V.: Remote sensing techniques for studying coastal ecosystems: an overview. J. Coastal Res., 27, 2-17, 2011. 38

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Knudby, A., Kenchington, E., and Murillo, F. J.: Modeling the distribution of Geodia sponges and sponge grounds in the Northwest Atlantic. PLOS One, 8, e82306, 2013.

Knudby, A., LeDrew, E., and Brenning, A.: Predictive mapping of reef fish species richness, diversity and biomass in Zanzibar using IKONOS imagery and machine-learning techniques. Remote Sens. Environ., 114, 1230-1241, 2010.

5 Konsberg Maritime: HISAS 1030 High resolution interferometric synthetic aperture sonar. URL http://www.km.kongsberg.com/ks/web/nokbg0397.nsf/AllWeb/86E9FFB43569CDEEC12576B9006D75C7/$file/HISAS_10 30_brochure_v1_lowres_v2.pdf, Accessed 3 December 2015.

Kostylev, V. E., Erlandsson, J., Ming, M. Y., and Williams, G. A.: The relative importance of habitat complexity and surface area in assessing biodiversity: fractal application on rocky shores. Ecol. Complex., 2, 272-286, 2005.

10 Krause, D. C., and Menard, H. W.: Depth distribution and bathymetric classification of some seafloor profiles. Mar. Geol., 3, 169-193, 1965.

Kukowski, N., Hampel, A., Hoth, S., and Bialas, J.: Morphotectonic and morphometric analysis of the Nazca plate and the adjacent offshore Peruvian continental slope - Implications for submarine landscape evolution. Mar. Geol., 254, 107-120, 2008.

15 Kunze, E., and Llewellyn Smith, S. G.: The role of small-scale topography in turbulent mixing of the global ocean. Oceanography, 17, 55-64, 2004.

Kwasnitschka, T., Hansteen, T. H., Devey, C. W., and Kutterolf, S.: Doing fieldwork on the seafloor: photogrammetric techniques to yield 3D visual models from ROV video. Comput. Geosci., 52, 218-226, 2013.

Lafon, V., Froidefond, J. M., Lahet, F., and Castaing, P.: SPOT shallow water bathymetry of a moderately turbid tidal inlet 20 based on field measurements. Remote Sens. Environ., 81, 136-148, 2002.

Lanier, A., Romsos, C., and Goldfinger, C.: Seafloor habitat mapping on the Oregon continental margin: A spatially nested GIS approach to mapping scale, mapping methods, and accuracy quantification. Mar. Geod., 30, 51-76, 2007.

Lawrence, E., Hayes, K. R., Lucieer, V. L., Nichol, S. L., Dambacher, J. M., Hill, N. A., Barrett, N., Kool, J., and Siwabessy, J.: Mapping habitats and developing baselines in offshore marine reserves with little prior knowledge: a critical evaluation of 25 a new approach. PloS One, 10, e0141051, 2015.

Lecours, V.: Terrain Attribute Selection for Spatial Ecology (TASSE), v.1.0, URL www.marinegis.com, 2015.

Lecours, V., and Devillers, R.: Assessing the spatial data quality paradox in the deep-sea, in Proceedings of spatial knowledge and information Canada 2015, 1-8, 2015.

39

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Lecours, V., Lucieer, V. L., Dolan, M. F. J., and Micallef, A.: An ocean of possibilities: applications and challenges of marine geomorphometry, in: Geomorphometry for geosciences, International Society for Geomorphometry, Poznań, Poland, 23-26, 2015a.

Lecours, V., Devillers, R., Schneider, D.C., Lucieer, V.L., Brown, C.J., and Edinger, E.N.: Spatial scale and geographic 5 context in benthic habitat mapping: review and future directions. Mar. Ecol. Prog. Ser., 535, 259-284, 2015b.

Lecours, V., Miles, L. L., Devillers, R., and Edinger, E. N.: Data analysis towards the multiscale characterization of cold- water coral and sponge habitats in Canadian waters. Tech. Rep. REQ. No. F6160-120010. Department of Fisheries and Oceans Canada, Newfoundland and Labrador Region, St. John’s, 2013.

Legleiter, C. J., Roberts, D. A., Marcus, W. A., and Fonstad, M. A.: Passive optical remote sensing of river channel 10 morphology and in-stream habitat: physical basis and feasibility. Remote Sens. Environ., 93, 493-510, 2004.

Leon, J. X., Phinn, S. R., Hamylton, S., and Saunders, M. I.: Filling the ‘white ribbon’ – a multisource seamless digital elevation model for Lizard Island, northern Great Barrier Reef. Int. J. Remote Sens., 34, 6337-6354, 2013.

Li, F. K., and Goldstein, R. M.: Studies of multibaseline spaceborne interferometric synthetic aperture radars. IEEE T. Geosci. Remote, 28, 88-97, 1990.

15 Li, H., and Wu, J.: Uncertainty analysis in ecological studies: an overview, in Scaling and uncertainty analysis in ecology: methods and applications, Springer, Dordrecht, The Netherlands, 45-66, 2006.

Li, K., Chen, J., Tarolli, P., Sofia, G., Feng, Z., and Li, J.: Geomorphometric multi-scale analysis for the automatic detection of linear structures on the lunar surface. Earth Sci. Front., 21, 212-222, 2014.

Li, Z., Zhu, Q., and Gold, C.: Digital terrain modeling: principles and methodology. CRC Press, United States, 319 p., 2005.

20 Liceaga-Correa, M. A., and Euan-Avila, J. I.: Assessment of coral reef bathymetric mapping using visible Landsat Thematic Mapper data. Int. J. Remote Sens., 23, 3-14, 2002.

Louchard, E. M., Reid, R. P., Stephens, F. C., Davis, C. O., Leathers, R. A., and Downes, T. V.: Optical remote sensing of benthic habitats and bathymetry in coastal environments at Lee Stocking Island, Bahamas: a comparative spectral classification approach. Limnol. Oceanogr., 48, 511-521, 2003.

25 Lucieer, V. L.: Object-oriented classification of sidescan sonar data for mapping benthic marine habitats. Int. J. Remote Sens., 29, 905-921, 2008.

Lucieer, V., and Lamarche, G.: Unsupervised fuzzy classification and object-based image analysis of multibeam data to map deep water substrates, Cook Strait, New Zealand. Cont. Shelf Res., 31, 1236-1247, 2011.

40

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Lucieer, V., Barrett, N., Hill, N., and Nichol, S. L.: Characterization of shallow inshore coastal reefs on the Tasman Peninsula, Southeastern Tasmania, Australia, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 481-492, 2012.

Lucieer, V., Hill, N. A., Barrett, N. S., and Nichol, S.: Do marine substrate ‘look’ and ‘sound’ the same? Supervised 5 classification of multibeam acoustic data using autonomous underwater vehicle images. Estuar. Coast. Shelf S., 117, 94-106, 2013.

Lundblad, E., Wright, D. J., Miller, J., Larkin, E. M., Rinehart, R., Naar, D. F., Donahue, B. T., Anderson, S. M., and Battista, T.: A Benthic Terrain Classification Scheme for American Samoa. Mar. Geod., 29, 89-111, 2006.

Lurton, X.: An introduction to underwater acoustics: principles and applications, 2nd edition, Springer/Praxis Publishing, 10 Berlin, Germany, 704 p., 2010.

Lyzenga, D. R.: Passive remote sensing techniques for mapping water depth and bottom features. Appl. Optics, 17, 379-383, 1978.

Lyzenga, D. R., Malinas, N. P., and Tanis, F. J.: Multispectral bathymetry using a simple physically based algorithm. IEEE T. Geosci. Remote, 44, 2251-2259, 2006.

15 Ma, S., Tao, Z., Yang, X., Yu, Y., Zhou, X., and Li, Z.: Bathymetry retrieval from hyperspectral remote sensing data in optical-shallow water. IEEE T. Geosci. Remote, 52, 1205-1212, 2014.

Ma, Z., Li, C., and Gao, X.: Accretion tectonics of the global seafloor and its evolution. Sci. China Ser. D, 41, 624-625, 1998.

Mackenzie, D.: Ocean floor is laid bare by new satellite data. Science, 277, 1921, 1997.

20 Macon, C., Wozencraft, J., Park, J. Y., and Tuell, G.: Seafloor and land cover classification through airborne lidar and hyperspectral data fusion. Int. Geosci. Remote Se., II-77-II-80, 2008.

Maritorena, S., Morel, A., and Gentili, B.: Diffuse-reflectance of oceanic shallow waters – influence of water depth and bottom albedo. Limnol. Oceanogr., 39, 1689-1703, 1994.

Mark, D. M.: Geomorphometric parameters: a review and evaluation. Geogr. Ann. A, 57, 165-177, 1975.

25 Martorelli, E., D’Angelo, S., Fiorentino, A., and Chiocci, F. L.: Nontropical carbonate shelf sedimentation: the archipelago Pontino (central Italy) case history, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 449-456, 2012.

McAdoo, B. G., Pratson, L. F., and Orange, D. L.: Submarine landslide geomorphology, US continental slope. Mar. Geol., 169, 103-136, 2000. 41

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

McArthur, M. A., Brooke, B. P., Przeslawski, R., Ryan, D. A., Lucieer, V. L., Nichol, S., McCallum, A. W., Mellin, C., Cresswell, I. D., and Radke, L. C.: On the use of abiotic surrogates to describe marine benthic biodiversity. Estuar. Coast. Shelf S., 88, 21-32, 2010.

McKenzie, D., and Bowin, C.: The relationship between bathymetry and gravity in the Atlantic Ocean. J. Geophys. Res., 81, 5 1903-1915, 1976.

McNutt, M.: The hunt for MH370. Science, 344, 947, 2014.

Metzger, E.G., and Hurlburt, H.E.: The importance of high horizontal resolution and accurate coastline geometry in modeling South China Sea Inflow. Geophys. Res. Lett., 28, 1059-1062, 2001.

Miall, A.: Architecture and sequence stratigraphy in Pleistocene fluvial systems in the Malay Basin, based on seismic time- 10 slice analysis. AAPG Bull., 86, 1201–1216, 2003.

Micallef, A., and Mountjoy, J. J.:. A topographic signature of a hydrodynamic origin for submarine gullies. Geology, 39, 115-118, 2011.

Micallef, A., Berndt, C., Masson, D. G., and Stow, D. A. V.: A technique for the morphological characterization of submarine landscapes as exemplified by debris flows of the Storegga Slide. J. Geophys. Res., 112, F02001, 2007a.

15 Micallef, A., Berndt, C., Masson, D. G., and Stow, D. A. V.: Scale invariant characteristics of the Storegga Slide and implications for large-scale submarine mass movements. Mar. Geol., 247, 46-60, 2008.

Micallef, A., Le Bas, T. P., Huvenne, V. A. I., Blondel, P., Hühnerbach, V., and Deidun, A.: A multi-method approach for benthic habitat mapping of shallow coastal areas with high resolution multibeam data. Cont. Shelf Res., 39-40, 14-26, 2012.

Micallef, A., Masson, D. G., Berndt, C., and Stow, D. A. V.: Morphology and mechanics of submarine spreading: A case 20 study from the Storegga Slide. J. Geophys. Res., 112, F03023, 2007b.

Micallef, A., Masson, D. G., Berndt, C., and Stow, D. A. V.: Development and mass movement processes of the north- eastern Storegga Slide. Quaternary Sci. Rev., 28, 433-448, 2009.

Micallef, A., Mountjoy, J., Barnes, P. M., Canals, M., and Lastras, G.: Geomorphic response of submarine canyons to tectonic activity: Insights from the Cook Strait canyon system, New Zealand. Geosphere, 10, 905-929, 2014a.

25 Micallef, A., Mountjoy, J. J., Canals, M., and Lastras, G.: Deep-seated bedrock landslides and submarine canyon evolution in an active tectonic margin: Cook Strait, New Zealand., in: Submarine Mass Movements and Their Consequences, Springer, London, United Kingdom, 201-212, 2012.

Micallef, A., Ribó, M., Canals, M., Puig, P., Lastras, G., and Tubau, X.: Space-for-time substitution and the evolution of a submarine canyon–channel system in a passive progradational margin. Geomorphology, 221, 34-50, 2014b. 42

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Minár, J., and Evans, I. S.: Elementary forms for land surface segmentation: the theoretical basis of terrain analysis and geomorphological mapping. Geomorphology, 95, 236-259, 2008.

Minár, J., Jenčo, M., Evans, I. S., Minár Jr., J., Kadlec, M., Krcho, J., Pacina, J., Burian, L., and Benová, A.: Third-order geomorphometric variables (derivatives): definition, computation and utilization of changes of curvatures. Int. J. Geogr. Inf. 5 Sci., 27, 1381-1402, 2013.

Mitchell, N. C.: Susceptibility of mid-ocean ridge volcanic islands and seamounts to large-scale landsliding. J. Geophys. Res., 108, 2397, 2003.

Mitchell, N. C.: Form of submarine erosion from confluences in Atlantic USA continental slope canyons. Am. J. Sci., 304, 590-611, 2004.

10 Mitchell, N. C.: Interpreting long-profiles of canyons in the USA Atlantic continental slope. Mar. Geol., 214, 75-99, 2005.

Mitchell, N. C., and Clarke, J. E. H.: Classification of seafloor geology using multibeam sonar data from the Scotian Shelf. Mar. Geol., 121, 143-160, 1994.

Mitchell, N. C., and Searle, R. C.: Fault scarp statistics at the Galapagos spreading centre from Deep Tow data. Mar. Geophys. Res., 20, 183-193, 1998.

15 Mitchell, N. C., and Somers, M. L.: Quantitative backscatter measurements with a long-range side-scan. IEEE J. Oceanic Eng., 14, 368-374, 1989.

Mitchell, N. C., Dade, W. B., and Masson, D. G.: Erosion of the submarine flanks of the Canary Islands. J. Geophys. Res., 108, 6002, 2003.

Mitchell, N. C., Masson, D. G., Watts, A. B., Gee, M. J. R., and Urgeles, R.: The morphology of the submarine flanks of 20 volcanic ocean islands: A comparative study of the Canary and Hawaiian hotspot islands. J. Volcanol. Geoth. Res., 115, 83- 107, 2002.

Mitchell, N. C., Tivey, M. A., and Gente, P.: Seafloor slopes at mid-ocean ridges from submersible observations and implications for interpreting geology from seafloor topography. Earth Planet. Sc. Lett., 183, 543-555, 2000.

Mobley, C. D., Sundman, L. K., Davis, C. O., Bowles, J. H., Downes, T. V., Leathers, R. A., Montes, M. J., Bissett, W. P., 25 Kohler, D. D. R., Reid, R. P., Louchard, E. M., and Gleason, A.: Interpretation of hyperspectral remote-sensing imagery by spectrum matching and look-up tables. Appl. Optics, 44, 3576-3592, 2005.

Moernaut, J., and De Batist, M.: Frontal emplacement and mobility of sublacustrine landslides: Results from morphometric and seismostratigraphic analysis. Mar. Geol., 285, 29-45, 2011.

43

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Mofield, H.O., Massell Symons, C., Londsdale, P., Gonzales, F.I., and Titov, V.V.: Tsunami scattering and earthquake faults in the deep Pacific Ocean. Oceanography, 17, 38-46, 2004.

Mohammadi, H., Rajabifard, A., and Williamson, I. P.: Enabling spatial data sharing through multi-source spatial data integration. Global Spatial Data Infrastructure Conference, Rotterdam, The Netherlands, 1-19, 2011.

5 Mohn, C., Rengstorf, A., White, M., Duineveld, G., Mienis, F., Soetaert, K., and Grehan, A.: Linking benthic hydrodynamics and cold-water coral occurrences: a high-resolution model study at three cold-water coral provinces in the NE Atlantic. Prog. Oceanog., 122, 92-104, 2014.

Monk, J., Ierodiaconou, D., Versace, V. L., Bellgrove, A., Harvey, E., Rattray, A., Laurenson, L., and Quinn, G. P.: Habitat suitability for marine fishes using presence-only modelling and multibeam sonar. Mar. Ecol. Prog. Ser., 420, 157-174, 2010.

10 Monk, J., Ierodiaconou, D., Bellgrove, A., Harvey, E., and Laurenson, L.: Remotely sensed hydroacoustics and observation data for predicting fish habitat suitability. Cont. Shelf Res., 31, S17-S27, 2011.

Mosalik, M., Błaszczyk, M., and Jania, J.: Statistical analysis of Brepollen bathymetry as a key to determine average depths on a glacier foreland. Geomorphology, 206, 262-270, 2014a.

Mosalik, M., Tegowski, J., Grabowiecki, P., and Zulichowska, M.: Principal Component and Cluster Analysis for 15 determining diversification of bottom morphology based on bathymetric profiles from Brepollen (Hornsund, Spitsbergen). Oceanologia, 56, 59-84, 2014b.

Mountjoy, J. J., McKean, J., Barnes, P. M., and Pettinga, J. R.: Terrestrial-style slow-moving earthflow kinematics in a submarine landslide complex. Mar. Geol., 267, 114-127, 2009.

Nayegandhi, M., Brock, J., and Wright, C.: Small-footprint, waveform-resolving lidar estimation of submerged and sub- 20 canopy topography in coastal environments. Int. J. Remote Sens., 30, 861-878, 2009.

Neidell, N. S.: Spectral studies of marine geophysical profiles. Geophysics, 31, 122-134, 1966.

NOAA (National Oceanic and Atmospheric Administration): Digital Coast, Office for Coastal Management. URL https://coast.noaa.gov/digitalcoast/, Accessed 24 February 2016.

Normile, D.: Lost at sea. Science, 344, 963-965, 2014.

25 Oksanen, J., and Sarjakoski, T.: Error propagation of DEM-based surface derivatives. Comput. Geosci., 31, 1015-1027, 2005.

Pacheco, A., Horta, J., Loureiro, C., and Ferreira, Ó.: Retrieval of nearshore bathymetry from Landsat 8 images: a tool for coastal monitoring in shallow waters. Remote Sens. Environ., 159, 102-116, 2015.

44

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Passalacqua, P., Belmont, P., Staley, D. M., Simley, J. D., Arrowsmith, J. R., Bode, C. A., Crosby, C., DeLong, S. B., Glenn, N. F., Kelly, S. A., Lague, D., Sangireddy, H., Schaffrath, K., Tarboton, D. G., Wasklewicz, T., and Wheaton, J. M.: Analyzing high resolution topography for advancing the understanding of mass and energy transfer through landscapes: a review. Earth-Sci. Rev., 148, 174-193, 2015.

5 Passaro, S., Barra, M., Saggiomo, R., Di Giacomo, S., Leotta, A., Uhlen, H., and Mazzola, S. : Multi-resolution morpho- bathymetric survey results at the Pozzuoli-Baia underwater archaeological site (Naples, Italy). J. Archaeol. Sci., 40, 1268- 1278, 2013.

Passaro, S., Ferranti, L., and De Alteris, G.: The use of high-resolution elevation histograms for mapping submerged terraces: Tests from the Eastern Tyrrhenian Sea and the Eastern Atlantic Ocean. Quatern. Int., 232, 238-249, 2011.

10 Passaro, S., Milano, G., D'Istanto, C., Ruggieri, S., Tonielli, R., Bruno, P. P., Sprovieri, M., and Marsella, E.: DTM-based morphometry of the Palinuro seamount (Eastern Tyrrhenian Sea): Geomorphological and volcanological implications. Geomorphology, 115, 129-140, 2010.

Peyton, L., and Boettcher, R.: Interpretation of incised valleys using new 3-D seismic techniques: A case history using spectral decomposition and coherency. The Leading Edge, 17, 1294–1296, 1998.

15 Pike, R. J., Evans, I. S., and Hengl, T.: Geomorphometry: a brief guide, in: Geomorphometry: concepts, software, applications, Elsevier, Amsterdam, The Netherlands, 3-30, 2009.

Pirtle, J. L., Weber, T. C., Wilson, C. D., and Rooper, C. N.: Assessment of trawlable and untrawlable seafloor using multibeam-derived metrics. Methods Oceanogr., 12, 18-35, 2015.

Pittman, S. J., Costa, B. M., and Battista, T. A.: Using lidar bathymetry and boosted regression trees to predict the diversity 20 and abundance of fish and corals. J. Coastal Res., 53, 27-38, 2009.

Pohl, C., and van Genderen, J. L.: Multisensor image fusion in remote sensing: concepts, methods and applications. Int. J. Remote Sens., 19, 823-854, 1998.

Posamentier, H. W.: Depositional elements associated with a basin floor channel–levee system: Case study from the Gulf of Mexico. Mar. Petrol. Geol., 20, 677–690, 2003.

25 Posamentier, H. W., and Kolla, V.: Seismic geomorphology and stratigraphy of depositional elements in deep-water settings. J. Sediment. Res., 73, 367–388, 2003.

Posamentier, H. W., Dorn, G. A., Cole, M. J., Beierle, C. W., and Ross, S. P.: Imaging elements of depositional systems with 3-D seismic data: a case study. SEPM, Gulf Coast Section, 17th Annual Research Conference, 213–228, 1996.

45

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Posamentier, H. W., Meizarwin, P. S., and Wisman, T.: Deep water depositional systems—ultra-deep Makassar Strait, Indonesia, in: Deep-water reservoirs of the world, SEPM, Gulf Coast Section, 20th Annual Research Conference, 806–816, 2000.

Pratson, L. F., and Ryan, W. B. F.: Automated drainage extraction in mapping the Monterey submarine drainage system, 5 California margin. Mar. Geophys. Res., 18, 757-777, 1996.

Preston, J. M., Christney, A. C., Bloomer, S. F., and Beaudet, I. L.: Seabed classification of multibeam sonar images. Oceans-IEEE, 2616-2623, 2001.

Purkis, S. J., Graham, N. A. J., and Riegl, B. M.: Predictability of reef fish abundance using remote sensing data in Diego Garcia (Chagos Archipelago). Coral Reefs, 27, 167-178, 2008.

10 Quadros, N. D., Collier, P. A., and Fraser, C. S.: Integration of bathymetric and topographic lidar: a preliminary investigation. Int. Arch. Photogrammetry, Remote Sens. Spat. Inf. Sci., 37, 315-320, 2008.

Ramillien, G., and Cazenave, A.: Global bathymetry derived from altimeter data of the ERS-1 geodetic mission. J. Geodyn., 23, 129-149, 1997.

Rattray, A., Ierodiaconnu, D., Monk, J., Versace, V. L., and Laurenson, L. J. B.: Detecting patterns of change in benthic 15 habitats by acoustic remote sensing. Mar. Ecol. Prog. Ser., 477, 1-13, 2013.

Rattray, A., Ierodiaconou, D., Monk, J., Laurenson, L. J. B., and Kennedy, P.: Quantification of spatial and thematic uncertainty in the application of underwater video for benthic habitat mapping. Mar. Geod., 37, 315-336, 2014.

Rengstorf, A. M., Grehan, A. J., Yesson, C., and Brown, C.: Towards high-resolution habitat suitability modelling of vulnerable marine ecosystems in the deep-sea: resolving terrain attribute dependencies. Mar. Geod., 35, 343-361, 2012.

20 Rengstorf, A. M., Yesson, C., Brown, C., and Grehan, A. J.: High-resolution habitat suitability modelling can improve conservation of vulnerable marine ecosystems in the deep sea. J. Biogeogr., 40, 1702-1714, 2013.

Rengstorf, A. M., Mohn, C., Brown, C., Wisz, M. S., and Grehan, A. J.: Predicting the distribution of deep-sea vulnerable marine ecosystems using high-resolution data: considerations and novel approaches. Deep Sea Res. Pt I, 93, 72-82, 2014.

Rigol-Sanchez, J. P., Stuart, N., and Pulido-Bosch, A.: ArcGeomorphometry: a toolbox for geomorphometric 25 characterisation of DEMs in the ArcGIS environment. Comput. Geosci., 85, 155-163, 2015.

Robinson, L. M., Elith, J., Hobday, A. J., Pearson, R. G., Kendall, B. E., Possingham, H. P., and Richardson, A. J.: Pushing the limits in marine species distribution modelling: lessons from the land present challenges and opportunities. Global Ecol. Biogeogr., 20, 211-226, 2011.

46

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Roman, C., and Singh, H.: Consistency based error evaluation for deep sea bathymetric mapping with robotic vehicles, in Proceedings of the 2006 IEEE international conference on robotics and automation, Orlando, FL, 3568-3574, 2006.

Ross, L. K., Ross, R. E., Stewart, H. A., and Howell, K. L.: The influence of data resolution on predicted distribution and estimates of extent of current protection of three ‘listed’ deep-sea habitats. PloS One, 10, e0140061, 2015.

5 Ross, R. E., and Howell, K. L.: Use of predictive habitat modelling to assess the distribution and extent of the current protection of ‘listed’ deep-sea habitats. Divers. Distrib., 19, 433-445, 2013.

Rovere, M., Gamberi, F., Mercorella, A., and Leidi, E.: Geomorphometry of a submarine mass-transport complex and relationships with active faults in a rapidly uplifting margin (Gioia Basin, NE Sicily margin). Mar. Geol., 356, 31-43, 2014.

Roy, S., Hovland, M., Noormets, R., and Olaussen, S.: Seepage in Isfjorden and its tributary fjords, West Spitsbergen. Mar. 10 Geol., 363, 146-159, 2015.

Samadzadeghan, F.: Data integration related to sensors, data and models. Int. Soc. Photogramm., Istanbul, Turkey, 2004.

Sandwell, D. T., and Smith, W. H. F.: Bathymetric estimation, in: Satellite altimetry and Earth Sciences: a handbook of techniques and applications, Academic Press, San Diego, United States, 441-458, 2001.

Sandwell, D. T., Gille, S. T., and Smith, W. H. F.: Bathymetry from space: oceanography, geophysics, and climate. 15 Geoscience Profesional Services, Bethesda, Maryland (U.S.), 2002.

Sappington, J. M., Longshore, K. M., and Thomson, D. B.: Quantifiying Landscape Ruggedness for Animal Habitat Anaysis: A case Study Using Bighorn Sheep in the Mojave Desert. J. Wildlife Manage., 71, 1419 -1426, 2007.

Schimel, A. C. G., Healy, T. R., Johnson, D., and Immenga, D.: Quantitative experimental comparison of single-beam, sidescan, and multibeam benthic habitat maps. ICES J. Mar. Sci., 67, 1766-1779, 2010a.

20 Schimel, A. C. G., Healy, T. R., McComb, P., and Immenga, D.: Comparison of a self-processed EM3000 multibeam echosounder dataset with a QTC View habitat mapping and a sidescan sonar imagey, Tamaki Strait, New Zealand. J. Coastal Res., 26, 714-725, 2010b.

Schimel, A. C., Ierodiaconou, D., Hulands, L., and Kennedy, D. M.: Accounting for uncertainty in volumes of seabed change measured with repeat multibeam sonar surveys. Cont. Shelf Res., 111, 52-68, 2015.

25 Serway, R. A., and Beichner, R. J.: Physics for scientists and engineers (Saunders golden sunburst series), 2nd edition, Saunders College Publishing, Philadelphia, United States, 1072 p., 1983.

Shary, P. A., Sharaya, L. S., and Mitusov, A. V.: The problem of scale-specific and scale-free approaches in geomorphometry. Geogr. Fis. Din. Quat., 28, 81-101, 2005.

47

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Shaw, J., Piper, D. J. W., Fader, G. B. J., King, E. L., Todd, B. J., Bell, T., Batterson, M. J., and Liverman, D. G. E.: A conceptual model of the deglaciation of Atlantic Canada. Quaternary Sci. Rev., 25, 2059-2081, 2006.

Shih, P. T.-Y., Chen, Y.-H., and Chen, J.-C.: Historic shipwreck study in Dongsha Atoll with bathymetric LiDAR. Archaeol. Prospect., 21, 139-146, 2014.

5 Smith, D. K., and Shaw, P. R.: Using topographic slope distributions to infer seafloor patterns. IEEE J. Oceanic Eng., 14, 338-347, 1989.

Smith, D. P., Kvitek, R.., Iampietro, P. J., and Wong, K.: Twenty-nine months of geomorphic change in upper Monterey Canyon (2002-2005). Mar. Geol., 236, 79-94, 2007.

Smith, M. J., Paron, P., and Griffiths, J. S.: Geomorphological mapping: methods and applications. Developments in Earth 10 Surface Processes Vol. 15, Elsevier Science, The Netherlands, 610 p., 2011.

Smith, W. H. F.: On the accuracy of digital bathymetric data. J. Geophys. Res., 98, 9591-9603, 1993.

Smith, W. H. F.: Seafloor tectonic fabric from satellite altimetry. Annu. Rev. Earth Planet. Sci., 26, 697-738, 1998.

Smith, W. H. F.: Introduction to this special issue on bathymetry from space. Oceanography, 17, 6-7, 2004.

Smith, W. H. F., and Marks, K. M.: Seafloor in the Malaysia Airlines flight MH370 search area. EOS. T. Am. Geophys, Un., 15 95, 173-180, 2014.

Smith, W. H. F., and Sandwell, D. T.: Bathymetry prediction from dense satellite altimetry and sparse shipboard bathymetry. J. Geophys. Res., 99, 12803-21824, 1994.

Smith, W. H. F., and Sandwell, D. T.: Global sea floor topography from satellite altimetry and ship depth soundings. Science, 277, 1956-1962, 1997.

20 Sofia, G., Pirotti, F., and Tarolli, P.: Variations in multiscale curvature distribution and signatures of LiDAR DTM errors. Earth Surf. Proc. Land., 38, 1116-1134, 2013.

Solan, M., Germano, J. D., Rhoads, D. C., Smith, C., Michaud, E., Parry, D., Wenzhöfer, F., Kennedy, B., Henriques, C., Battle, E., Carey, D., Iocco, L., Valente, R., Watson, J., and Rosenberg, R.: Towards a greater understanding of pattern, scale and process in marine benthic systems: a picture is worth a thousand worms. J. Exp. Mar. Biol. Ecol., 285-286, 313-338, 25 2003.

Solsten, B. L., and Aitken, A. E.: An application of GIS techniques to assess the risk of disturbance of archaeological sites by mass movement and marine flooding in Auyuittuq National Park Reserve, Nunavut. Geogr. Phys. Quatern., 60, 81-92, 2006.

48

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Speight, J. G.: A parametric approach to landform regions, in: Progress in geomorphology: papers in honour of David L. Linton, Alden Press, London, United Kingdom, 239-250, 1974.

Stepinski, T., and Vilalta, R.: Digital topography models for Martian surfaces. IEEE T. Geosci. Remote, 2, 260-264, 2005.

Stojic, M., Chandler, J., Ashmore, P., and Luce, J.: The assessment of sediment transport rates by automated digital 5 photogrammetry. Photogramm. Eng. Rem. S., 64, 387-395, 1998.

Stoker, J. M., Tyler, D. J., Turnipseed, D. P., Van Wilson Jr., K., and Oimoen, M. J.: Integrating disparate lidar datasets for a regional storm tide inundation analysis of hurricane Katrina. J. Coastal Res., 53, 66-72, 2009.

Storlazzi, C. D., Fregoso, T. A., Figurski, J. D., Freiwald, J., Lonhart, S. I., and Finlayson, D. P.: Burial and exhumation of temperate bedrock reefs as elucidated by repetitive high-resolution sea floor sonar surveys: spatial patterns and impacts to 10 species’ richness and diversity. Cont. Shelf Res., 55, 40-51, 2013.

Stow, D.A.V., Hernandez-Molina, F.J., Llave, E., Sayago-Gil, M., Diaz del Rio, V., and Branson, A.: Bedform-velocity matrix: The estimation of bottom current velocity from bedform observations. Geology, 37, 327-330, 2009.

Stretch, R. C., Mitchell, N. C., and Portaro, R. A.: A morphometric analysis of the submarine volcanic ridge south-east of Pico Island, Azores. J. Volcanol. Geoth. Res., 156, 35-54, 2006.

15 Stumpf, R. P., Holderied, K., and Sinclair, M.: Determination of water depth with high-resolution satellite imagery over variable bottom types. Limnol. Oceanogr., 48, 547-556, 2003.

Su, H., Liu, H., Wang, L., Filippi, A. M., Heyman, W. D., and Beck, R. A.: Geographically adaptive inversion model for improving bathymetric retrieval from satellite multispectral imagery. IEEE T. Geosci. Remote, 52, 465-476, 2014.

Teide Group: Morphometric interpretation of the northwest and southeast slopes of Tenerife, Canary Islands. J. Geophys. 20 Res., 102, 20325-20342, 1997.

Tempera, F., Giacomello, E., Mitchell, N. C., Campos, A. S., Braga Henriques, A., Bashmachnikov, I., Martins, A., Mendonça, A., Morato, T., Colaço, A., Porteiro, F. M., Catarino, D., Gonçalves, J., Pinho, M. R., Isidro, E. J., Santos R. S., and Menezes, G.: Mapping the Condor seamount seafloor environment and associated biological assemblages (Azores, NE Atlantic), in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic 25 habitats, Elsevier, Amsterdam, The Netherlands, 807-818, 2012.

Thiers, L., Louzao, M., Ridoux, V., Corre, M. L., Jaquemet, S., and Weimerskirch, H.: Combining methods to describe important marine habitats for top predators: applications to identify biological hotspots in tropical waters. PLoS ONE, 9, e115057, 2014.

49

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Thorsnes, T., Bjarnadóttir, L. R., Lågstad, P., Olsen, K.T., Holte, B., Dolan, M.F.J., Buhl-Mortensen, P. and Hodnesdal, H.: Assessment of AUV-borne acoustic and visual tools for seabed mapping - cruise report. NGU Report 2016.006. Geological Survey of Norway, Trondheim. In press.

Tong, R., Purser, A., Guinan, J., and Unnithan, V.: Modeling the habitat suitability of deep-water gorgonian corals based on 5 terrain variables. Ecol. Inform., 13, 123-132, 2013.

Usery, E. L., Finn, M. P., and Starbuck, M.: Integrating data layers to support the national map of the United States. Int. Cartogr. Conf., Corua, Spain, 2005.

Vachtman, D., Mitchell, N. C., and Gawthorpe, B.: Morphologic signatures in submarine canyons and gullies, central USA Atlantic continental margins. Mar. Petrol. Geol., 41, 250-263, 2013.

10 van Lancker, V., Moerkerke, G., Du Four, I., Verfaillie, E., Rabaut, M., and Degraer, S.: Fine-scale geomorphological mapping of sandbank environments for the prediction of macrobenthic occurrences, Belgian part of the North Sea, in: Seafloor geomorphology as benthic habitat: GeoHab atlas of seafloor geomorphic features and benthic habitats, Elsevier, Amsterdam, The Netherlands, 251-260, 2012.

Van Rein, H. B., Brown, C. J., Quinn, R., and Breen, J.: A review of sublittoral monitoring methods in temperate waters: a 15 focus on spatial scale. Underwater Technol., 28, 99-113, 2009.

Vierod, A. D. T., Guinotte, J. M., and Davies, A. J.: Predicting the distribution of vulnerable marine ecosystems in the deep sea using presence-background models. Deep-Sea Res. II, 99, 6-18, 2014.

Wang, C. K., and Philpot, W. D.: Using airborne bathymetric LiDAR to detect bottom type variation in shallow waters. Remote Sens. Environ., 106, 123-135, 2007.

20 Watts, A. B.: On geoid heights derived from the GEOS 3 altimeter data along the Hawaiian-Emperor seamount chain. J. Geophys. Res., 84, 3817-3826, 1979.

Watts, A. B., and Ribe, N. M.: On geoid heights and flexure of the lithosphere at seamounts. J. Geophys. Res., 89, 11152- 11170, 1984.

Wedding, L. M., Friedlander, A. M., McGranaghan, M., Yost, R. S., and Monaco, M. E.: Using bathymetric LiDAR to 25 define nearshore benthic habitat complexity: implications for management of reef fish assemblages in Hawaii. Remote Sens. Environ., 112, 4159-4165, 2008.

Weiss, A. D.: Topographic Position and Landform Analysis (poster), ESRI User Conference, San Diego, USA, 2001.

Westaway, S. J., Lane, S. N., and Hicks, D. M.: Remote survey of large-scale braided rivers using digital photogrammetry and image analysis. Int. J. Remote Sens., 24, 795-816, 2003. 50

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

White, M., Mohn, C., and Orren, M. J.: Physical processes and seamount productivity, in: Seamounts: ecology, fisheries and conservation, Blackwell Publishing, United Kingdom, 65-84, 2007.

Whittingham, M. J., Percival, S. M., and Brown, A. F.: Nest-site selection by golden plover: why do shorebirds avoid nesting on slopes? J. Avian Biol., 33, 184-190, 2002.

5 Winterbottom, S. J., and Gilvear, D. J.: Quantification of channel bed morphology in gravel-bed rivers using airborne multispectral imagery and aerial photography. Regul. River., 13, 489-499, 1997.

Wilson, M. F. J.: Deep sea habitat mapping using a Remotely Operated Vehicle: mapping and modelling seabed terrain and benthic habitat at multiple scales in the Porcupine Seabight, SW Ireland. PhD Thesis. National University of Ireland, Galway, 2006.

10 Wilson, M. F. J., O’Connell, B., Brown, C., Guinan, J. C., and Grehan, A. J.: Multiscale Terrain Analysis of Multibeam Bathymetry Data for Habitat Mapping on the Continental Slope. Mar. Geod., 30, 3-35, 2007.

Wise, S. M.: The effect of GIS interpolation errors on the use of digital elevation models in geomorphology, in: Landform monitoring, modelling and analysis, Wiley, Chichester, England, 139-164, 1998.

Wood, J. 1996. The Geomorphological Characterisation of Digital Elevation Models. PhD Thesis. University of Leicester.

15 Wood, J. 2009. Landserf Version 2.3 (www.landserf.org).

Wood, L. J.: Quantitative seismic geomorphology and reservoir architecture of clastic depositional systems: the future of uncertainty analysis in exploration and production. AAPG Ann. Meet., 182–183, 2003.

Wood, L. J., and Dragicevic, S.: GIS-based multicriteria evaluation and fuzzy sets to identify priority sites for marine protection. Biodivers. Conserv., 16, 2539-2558, 2007.

20 Wright, D. J., Pendleton, M., Boulware, J., Walbridge, S., Gerlt, B., Eslinger, D., Sampson, D., and Huntley, E.: ArcGIS Benthic Terrain Modeler (BTM), v. 3.0, Environmental Systems Research Institute, NOAA Coastal Services Center, Massachusetts Office of Coastal Zone Management, 2012.

Yang, F., Li, J., Wu, Z., Jin, X., Chu, F., and Kang, Z.: A post-processing method for the removal of refraction artifacts in multibeam bathymetry data. Mar. Geod., 30, 235-247, 2007.

25 Zavalas, R., Ierodiaconou, D., Ryan, D., Rattray, A., and Monk, J.: Habitat classification of temperate marine macroalgal communities using bathymetric LiDAR. Remote Sens., 6, 2154-2175, 2014.

Zeng, H. L., and Hentz, T. F.: High-frequency sequence stratigraphy from seismic sedimentology applied to Miocene, Vermilion Block 50, Tiger Shoal area, offshore Louisiana. AAPG Bull., 88, 153–174, 2004.

51

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Zevenbergen, L. W., and Thorne, C.: Quantitative analysis of land surface topography. Earth Surf. Proc. Land., 12, 47-56, 1987.

Zhang, J.: Multi-source remote sensing data fusion: status and trends. Int. J. Image Data Fusion, 1, 5-24, 2010.

Zhou, Q., and Liu, X.: Analysis of errors of derived slope and aspect related to DEM data properties. Comput. Geosci., 30, 5 369-378, 2004.

Zieger, S., Stieglitz, T., and Kininmonth, S.: Mapping reef features from multibeam sonar data using multiscale morphometric analysis. Mar. Geol., 264, 209-217, 2009.

Zuber, M. T., Smith, D. E., Phillips, R. J., Solomon, S. C., Neumann, G. A., Hauck II, S. A., Peale, S. J., Barnouin, O. S., Head, J. W., Johnson, C. L., Lemoine, F. G., Mazarico, E., Sun, X., Torrence, M. H., Freed, A. M., Klimczak, C., Margot, J.- 10 L., Obserst, J., Perry, M. E., McNutt Jr., R. L., Balcerski, J. A., Michel, N., Talpe, M. J., and Yang, D.: Topography of the northern hemisphere of Mercury from MESSENGER laser altimetry. Science, 336, 217-220, 2012

52

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

1 Tables

2 Table 1: Summary of the most commonly used terrain attributes in marine-based studies, as well as an indication of some common calculation approaches. Modified after 3 Dolan et al. (2012). The term ‘multiple scale’ refers to terrain attributes derived in turn using analysis windows of different sizes. ‘Multiscale’ refers to indices derived 4 simultaneously over a range of window sizes. The more general term ‘multi-scale’ is used in this paper to refer to both types of analysis as well as geomorphometric 5 analysis using data of different resolutions.

Slope Orientation Curvature Terrain Variability Ecological Stability of sediments (ability to Degree of exposure to Index of exposure/shelter e.g. on a peak or Index of degree of habitat relevance live in/on sediments) dominant and/or local in a hollow (food supply, sedimentation, structure, shelter from currents from a particular predators etc.) exposure/predators (link to life Local acceleration of currents direction (food supply, stages). (food supply, exposure, etc.). sedimentation, larval Structural diversity linked to dispersion etc.) biodiversity Stability of sediments (grain size). Relation to direction of Flow, channelling of sediments/currents, Terrain variability and dominant geomorphic hydrological and glacial processes. structures present reflect Geomorphological Local acceleration of currents processes. dominant geomorphic relevance (erosion, movement of sediments, Useful in the classification of landforms processes. creation of bedforms). Slope (Lundblad et al., 2006; Aspect (Galparsoro et al., Mean curvature (Dolan et al., 2008) Rugosity (Dunn and Halpin, Lanier et al., 2007; Micallef et 2009), northness and 2009) Commonly Profile curvature (Guinan et al., 2009)) al., 2012; Dolan and Lucieer, eastness (Monk et al., 2011) computed terrain Vector Ruggedness Measure 2014) Plan curvature (Ross et al., 2015) attribute and (VRM) (Tempera et al., 2012) example marine- Bathymetric Position Index (BPI) (Monk et Relative Relief (Elvenes, 2013) based reference al., 2010; Pirtle et al., 2015) Fractal Dimension (Wilson et al., 2007)

53

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Commonly used Single scale: Single scale: Single scale: Single scale: terrain attribute Slope: ArcGIS Spatial Analyst Aspect: ArcGIS Spatial Mean, Profile and Plan curvature ArcGIS Rugosity (surface area/planar and software (Horn, 1981) Analyst (Horn, 1981) Spatial Analyst (Zevenbergen and Thorne, area ratio) (Jenness, 2004) (algorithm 1987) reference) Statistical Aspect (northness/eastness) BTM toolbox (Wright et al., 2012)

Multi-scale: Multi-scale: Multi-scale: Multi-scale: Multiple scale slope: Multiple scale aspect: Several measures of multiple scale Multiple scale VRM r.param.scale, Landserf (Wood, r.param.scale, Landserf curvature: r.param.scale, Landserf (Wood, (Sappington et al, 2007) 1996) (Wood, 1996) 1996) Multiple scale relative relief Multiscale slope: Landserf Multiscale aspect: Landserf Multiple scale BPI (Lundblad et al., 2006) (Erikstad et al., 2013 and (Wood, 1996) (Wood, 1996) references therein) Multiscale curvature Landserf (Wood, 1996) Multiple and multiscale fractal dimension - Landserf (Wood,

1996)

54

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Table 2: List of studies published in the last three decades that applied geomorphometric techniques to submarine landscapes.

Technique Reference General geomorphometry Morphometric attributes/basic geometrical analysis (Adams and Schlager, 2000) (De Moustier and Matsumoto, 1993) (Teide Group, 1997) Morphometric attributes and their statistical analyses (Berkson and Matthews, 1983) (Booth and O'Leary, 1991) (Chakraborty et al., 2001) (Goff and Jordan, 1988) (Kukowski et al., 2008) (Micallef et al., 2007a) (Mitchell et al., 2000) (Mosalik et al., 2014a) (Passaro et al., 2010) (Passaro et al., 2011) (Smith and Shaw, 1989) Spectral analysis (Fox and Hayes, 1985) (Fox, 1996) (Gilbert and Malinverno, 1988) (Goff and Tucholke, 1997) (Mosalik et al., 2014b) Geostatistical methods (Herzfeld, 1989) (Herzfeld and Higginson, 1996) (Ismail et al., 2015) Feature-based quantitative representation (Harrison et al., 2011) (Micallef et al., 2007a) (Micallef et al., 2007b) (Mitchell and Clarke, 1994) (Pratson and Ryan, 1996) Neural networks (Jiang et al., 1993) Other techniques (Mountjoy et al., 2009) (Preston et al., 2001) Specific geomorphometry (Casalbore et al., 2011) (Gee et al., 2001)

55

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

(Green and Uken, 2008) (Haflidason et al., 2005) (Hühnerbach et al., 2004) (Issler et al., 2005) (McAdoo et al., 2000) (Micallef et al., 2008) (Micallef and Mountjoy, 2011) (Micallef et al., 2012) (Micallef et al., 2014b) (Micallef et al., 2014a) (Mitchell and Searle, 1998) (Mitchell et al., 2002) (Mitchell et al., 2003) (Mitchell, 2003) (Mitchell, 2004) (Mitchell, 2005) (Moernaut and De Batist, 2011) (Rovere et al., 2014) (Roy et al., 2015) (Stretch et al., 2006) (Vachtman et al., 2013)

56

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figures

Fig. 1: Cumulative number of publications (articles or reviews) listed in the Scopus database mentioning specific keywords (see legends) in their title, abstracts of keywords, by the end of 2015. For the applications, the keywords “topographic variable*”, “topographic attribute*”, “terrain attribute*” or “terrain variable*” were used, without “marine” (top) or with “marine” (bottom).

Fig. 2: According to Pike et al. (2009), geomorphometry is commonly implemented in five steps, here adapted to the application of geomorphometry to the marine environment (left). The panels on the right describe the structure and elements addressed in Sect. 2 to Sect. 6 of this paper.

Fig. 3: Example of elements that can be extracted and visualised when using the CUBE algorithm, using the ROV-based dataset from Fig. 4 (source: Lecours and Devillers, 2015). On top, the components contributing to the horizontal and vertical TPUs can be studied. Other marginal contribution to the vertical TPU include roll and pitch of the platform, timing of the inertial measurement unit, and uncertainty associated with the sonar system (range and angle). The combination of the GPS and delta draft provides the three-dimensional position of the soundings (x, y, z); in ROV-based research, the positional accuracy decreases with depth (Lecours and Devillers, 2015). On the bottom, it is possible to visualise how the uncertainty and the density of soundings vary spatially. Other spatial information that CUBE provides include the standard deviation on estimated depth values based on the hypotheses generated by CUBE, and information on the hypotheses count and strength.

Fig. 4: Examples of errors and artefacts found in different datasets and their impact on derived terrain attributes. The top panels represent data from GEBCO (2014), which uses radar altimetry to fill in the gaps between higher-resolution, freely available bathymetric data. The main artefacts that can be observed are caused by the interpolation method that was used to combine the different datasets. For instance, a linear artefact following a Southwest to Northeast axis can be observed as a result of the combination of one SBES acoustic survey line with radar altimetry data. Similarly, some “spots” can be seen in the middle of the panels (South to North direction). These artefacts, especially apparent in the curvature, are caused by the merging of punctual lead line measurements with the radar altimetry data. Finally, a slight gridding artefact can be observed in the curvature (i.e. thin vertical and horizontal linear features). The middle panels show ship-based MBES data (Brown et al., 2012). The obvious artefacts follow the surveying pattern of the vessel, and are mainly caused by the roll motion of the vessel during the survey. Finally, the bottom row of panels corresponds to ROV-based MBES collected from 20 m above the seafloor in the deep sea (Lecours et al., 2013). In this case, the artefacts are caused by a combination of heave and other platform motions; the ancillary data collected to account for this motion are too uncertain at this depth to appropriately correct for the errors (Lecours and Devillers, 2015). The “spots” observed in the bottom and top of the derived terrain attributes are spurious soundings that can be removed in bathymetric software during post-processing. Note differences in spatial resolutions (left axis) and cartographic scales. Depth values of the top left panel range from 60 to 4,275 m deep, those of the middle left panel range from 20 to 105 m deep, and those of the bottom left panel range from 2,345 to 2,425 m deep. Lighter blue is shallower.

57

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Fig. 5: Illustration of the main types of terrain attributes that can be derived from bathymetry data. Modified after Wilson et al. (2007).

Fig. 6: Proportion of the marine geomorphometry articles or reviews compiled in Fig. 1 that corresponds to each of the types of applications presented in Sect. 6.

Fig. 7: Indicative workflow showing the use of terrain attributes in predictive habitat mapping. Generally following some pre-selection of variables the observed habitat points (response variable) are combined with full coverage predictor variables selected from bathymetry, terrain attributes and other environmental variables as available to form the input to a habitat model which will be used to predict a full-coverage habitat map. The choice of habitat model will depend on the study in question but is typically either a statistical (e.g. GLM) or machine-learning based model (e.g. Random Forest). Observed habitat points are classified from visual or physical samples of the seabed. Terrain attributes are typically multi-scale and may include general and/or specific geomorphometry. Other environmental variables may include, for example, oceanographic data (temperature, salinity, current speed etc.), geological data (e.g. grain size). A similar workflow applies to modelling of single species or communities, where the output will be a continuous map indicating the probability of occurrence within the study area, rather than a categorical map as shown here.

Fig. 8: Example of the use of marine geomorphometry to semi-automatically map the components of mega-scale submarine landslide offshore Norway (adapted from Micallef et al. (2009)). Figures a-c show the trough depth, ridge length and ridge spacing extracted from a multibeam echosounder map of the north-eastern Storegga Slide using ridge characterisation techniques (Micallef et al., 2007b). Figure d is a classification map generated by using these ridge characteristic maps as input layers in an unsupervised clustering algorithm (ISODATA). Figure e is an interpretative map of the range of spreading events based on figures a-d. Other mass movements and geological processes and structures have been interpreted using geomorphometric mapping (Micallef et al., 2007a).

58

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 1

59

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 2

60

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 3

61

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 4

62

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 5

63

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 6

64

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 7

65

Hydrol. Earth Syst. Sci. Discuss., doi:10.5194/hess-2016-73, 2016 Manuscript under review for journal Hydrol. Earth Syst. Sci. Published: 1 March 2016 c Author(s) 2016. CC-BY 3.0 License.

Figure 8

66